Aaron
Kleine‡
a,
Cem L.
Altan‡
bc,
U. Ecem
Yarar
b,
Nico A. J. M.
Sommerdijk
c,
Seyda
Bucak
*b and
Simon J.
Holder
*a
aDepartment of Chemical Engineering, Yeditepe University, Istanbul, 34755, Turkey. E-mail: seyda@yeditepe.edu.tr
bLaboratory of Materials and Interface Chemistry, and Soft Matter CryoTEM Research Unit, Department of Chemical Engineering and Chemistry, Eindhoven University of Technology, PO Box 513, 5600 MB Eindhoven, The Netherlands. E-mail: N.Sommerdijk@tue.nl
cFunctional Materials Group, School of Physical Sciences, University of Kent, Canterbury, Kent CT2 8EN, UK. E-mail: S.J.Holder@kent.ac.uk
First published on 10th September 2013
A facile synthetic route to poly(ethylene imine)-graft-poly(oligo(ethylene glycol methyl ether)) (PEI-graft-POEGMA) functionalised superparamagnetic magnetite nanoparticles is described. The polymerisation of OEGMA from a model molecular amide demonstrated the feasibility of POEGMA synthesis under mild ATRP conditions (20 °C in ethanol) albeit with low initiator efficiencies. DFT studies suggest that the amide functionality is intrinsically of lower activity than ester functional monomers and initiators for atom transfer polymerisation (ATRP) as a consequence of higher bond dissociation energies and bond dissociation free energies (BDFE). However these studies further highlighted that use of an appropriate solvent could reduce the free energy of dissociation thereby reducing the relative difference in BDFE between the ester and amide groups. A commercial branched PEI sample was functionalised by reaction with 2-bromo-2-methylpropanoyl bromide giving an amide macroinitiator suitable for the atom transfer radical polymerisation (ATRP) of oligo(ethylene glycol methyl ether) methacrylate. The resulting PEI-graft-POEGMA copolymers were characterised by SEC, FT-IR and 1H and 13C NMR spectroscopy. PEI-graft-POEGMA coated magnetite nanoparticles were synthesised by a basic aqueous co-precipitation method and were characterised by transmission electron microscopy, thermogravimetric analysis and vibrating sample magnetometry and dynamic light scattering. These copolymer coated magnetite nanoparticles were demonstrated to be effectively stabilised in an aqueous medium. Overall the particle sizes and magnetic and physical properties of the coated samples were similar to those of uncoated samples.
Magnetite (Fe3O4), which contains both Fe+2 and Fe+3 ions is the iron oxide type most used as the basis for magnetic nanoparticles due to its biocompatibility and excellent magnetic properties. If the grain size of magnetite particles is less than about 12 nm, then individual particles can have single magnetic domains leading to superparamagnetic behavior but suitable stabilization of the particles is needed in order to prevent aggregation, while keeping the superparamagnetic properties unchanged. Many synthesis methods have been reported on the aqueous synthesis of superparamagnetic magnetite nanoparticles of which chemical co-precipitation is the most frequently used owing to its simplicity. In this method Fe+2 and Fe+3 ions are mixed in a stoichiometric ratio of 1 to 2 in the presence of a base at high pH under inert conditions. The latter is to prevent oxidation of particles to maghemite which is also a superparamagnetic iron oxide but has a lower saturation magnetization value. The bare magnetite nanoparticles obtained are generally stabilized with surfactants or polymers by surface functionalization, preventing sedimentation and/or aggregation in solution.2 It has previously been shown that poly(ethylene imine) (PEI) can adsorb onto magnetite nanoparticles as a primary layer and a secondary layer of poly(ethylene oxide)-co-poly(glutamic acid) can give particles long term stability in physiological salt solution.9 Poly(ethylene glycol) (PEG) or poly(ethylene oxide) (PEO) has also been used by several groups to obtain stability in aqueous media.10–12 PEO possesses the advantage of being non-toxic and biocompatible and is widely utilized in biomedical applications. In contrast PEI has been investigated as an gene therapy carrier due to the high number of ammonium groups that are available to electrostatically interact with phosphate groups in DNA.13–15 Such polymeric ammonium groups are ideal cationic stabilisers for the magnetite water interface however PEI is relatively cytotoxic. In this study a PEI-graft-POEGMA polymer was designed as a stabilizer for magnetite to take advantage of the adsorption ability of the PEI backbone with the magnetite surface and the steric stabilisation from the comb-like of POEGMA chains.16 Copolymers formed from PEI and stimulus responsive polymers have been used to create systems which are ideal for controlled release.17,18 The PEI-graft-POEGMA copolymers presented in this research were specifically designed to take advantage of a number of design principles for nanoparticle stabilisation and biomedical application. The PEI backbone (of which a significant variety is commercially available) fulfils three functions, (1) to provide a number of amine groups that can directly interact with the iron oxide surface, (2) to provide amine groups amenable to functionalisation via simple and well-developed amide synthesis techniques, (3) provide the ability to fine tune stabiliser properties through the use of PEI of varying molecular architectures (branched and linear) and various molecular weights. For this initial study a low molecular weight branched PEI sample was chosen. Recent research reported on improved nanoparticle stability through the use of dendritic stabilisers relative to linear analogues.19 Whilst dendritic architectures are attractive for numerous applications the synthesis of such compounds is multi-step, time-consuming and costly. In contrast branched and hyperbranched architectures offer many of the advantages at a fraction of the cost and synthetic effort. The POEGMA polymer component is a bottle-brush type polymer with a polymethacrylic backbone with side chains of short chains of ethylene oxide. POEGMA has shown significant promise for bioconjugate systems due to its water solubility, protein and cell resistance and thermoresponsive properties.20–23 Thermoresponsive polymer coatings on magnetic nanoparticles to date have typically been introduced after particle synthesis by graft polymerisation or by seeded precipitation polymerisation and the most commonly employed thermoresponsive polymers to date has been poly(n-iso-propylacrylamide) and POEGMA.7,18,21,22,24–34 POEGMA is a poly methacrylate with graft oligo(ethylene oxide) side chains that has many of the properties of PEO whilst remaining amorphous and whose LCST can be fine tuned by varying side-chain length, end-groups and co-polymerising with other monomers.
PEI-graft-POEGMA based materials which can have their thermoresponsive behaviour tuned by the copolymerisation of 2-(2-methoxyethoxy)ethyl methacrylate (MEO2MA) and OEGMA have previously been synthesised by polymerising the copolymer first, then coupling it to the PEI core.35 Whilst this technique produced desirable materials (for gene delivery), it relied on a multiple stage synthesis using harsh conditions. Atom transfer radical polymerisation (ATRP) has been shown to be a robust and versatile technique for the creation of a wide range of polymers in relatively mild conditions.36–41 In this paper we will demonstrate that (i) OEGMA can be polymerised using a molecular amide initiator, (ii) the low initiator efficiency of amides in ATRP is a consequence of higher bond dissociation energies of the C–Br bond in amides, (iii) PEI can be modified to give a macro-initiator for ATRP by the reaction of amines within the PEI with an ATRP initiating moiety and (iv) the polymerisation of OEGMA from this modified PEI initiator is possible. This approach allows for the immediate and direct incorporation of the thermoresponsive functionality into the actual stabiliser structure with no need for post-functionalisation of the particles. Furthermore we will demonstrate the application of the resultant copolymer as a stabiliser for the synthesis of aqueous dispersed magnetite nanoparticles to obtain stable magnetic nano-materials in aqueous suspension.
Particle sizes, distributions and morphologies of the nanoparticles were analyzed by FEI Tecnai G2 Sphera transmission electron microscope (TEM) operating at 200 kV by drying 30 μl of samples on carbon coated 200 mesh copper grids. Phase identification of synthesized nanoparticles was obtained by Rigaku X-ray diffractometer (XRD) by scanning 2-theta range of 20° to 70° at room temperature with 0.02 theta increments per 10 s. Magnetic properties of both bare and PEI-b-POEGMA coated particles were analyzed by vibrating sample magnetometer at dry state and room temperature.
DLS measurements were performed on the PEI-graft-POEGMA stabilised nanoparticle dispersions using a Malvern Zetasizer Nano ZS with Dispersion Technology Software (DTS) version 5.0 software. All measurements of 10 scans were repeated three times and the average at each temperature reported.
2,2′-dinonyl-4,4′-dipyridine was used as the copper ligand and ethanol as the solvent with all reactions carried out at 25 °C. Representative results are given in Table 1. In agreement with previous reports of ATRP using amide ligands, the use of Cu(I)Br catalyst led to poorly controlled reactions with low conversions; replacing the Cu(I)Br with Cu(I)Cl led to relatively narrower polydispersity products but with poor initiation efficiencies. The effect of the Cu(I)Cl is to reduce the relative rate of propagation (of the methacrylate) to initiation (of the amide initiator). Increasing the total amount of solvent gave products with Mw/Mn < 1.3 and initiator efficiencies of around 0.7 to 0.8.
I | [M]![]() ![]() |
X | t (h) | M n(exp) | M n(th) | M w/Mn | Conv. (%) | I eff |
---|---|---|---|---|---|---|---|---|
a Solvent![]() ![]() ![]() ![]() |
||||||||
EBriB | 50 | Br | 24 | 11![]() |
12![]() |
1.31 | 81 | 1.02 |
EBriB | 100 | Br | 24 | 23![]() |
25![]() |
1.29 | 84 | 1.05 |
PBriBA | 50 | Br | 34 | 2750 | 4050 | 1.30 | 27 | 1.47 |
PBriBA | 50 | Br | 42 | 4300 | 3150 | 1.66 | 21 | 1.33 |
EBriB | 50 | Cl | 24 | 11![]() |
11![]() |
1.26 | 76 | 1.03 |
PBriBA | 50 | Cl | 20 | 19![]() |
6000 | 1.31 | 40 | 0.30 |
PBriBA | 50 | Cl | 28 | 20![]() |
7950 | 1.33 | 53 | 0.39 |
PBriBA | 50 | Cl | 36 | 18![]() |
8900 | 1.18 | 59 | 0.48 |
PBriBA | 50 | Cl | 48 | 21![]() |
11![]() |
1.23 | 74 | 0.51 |
EBriBa | 50 | Cl | 20 | 12![]() |
12![]() |
1.17 | 80 | 0.95 |
EBriBa | 50 | Cl | 67 | 14![]() |
13![]() |
1.30 | 93 | 0.94 |
PBriBAa | 50 | Cl | 48 | 13![]() |
9400 | 1.26 | 63 | 0.69 |
PBriBAa | 50 | Cl | 66 | 12![]() |
9800 | 1.13 | 65 | 0.79 |
DFT calculations were carried out on methyl 2-bromo-2-methylpropanoate (MBriP), N-methyl 2-bromo-2-methylpropanamide (MBriPA), N,N-dimethyl 2-bromo-2-methylpropanamide (MBriPA2), methyl 2-chloro-2-methylpropanoate (MCliP) and N-methyl 2-chloro-2-methylpropanamide (MCliPA) (Scheme 3). Geometries were optimised using the B3LYP functional with the 6-31+Gd basis set previously employed for studies on ATRP initiators.63–66 Since the B3LYP functional is known to give very inaccurate values for thermochemical calculations and in particular free bond dissociation energies, further functionals were employed for free energy calculations.67,68 For the single point energy calculations we used the BMK and M06-2X functionals both of which have been reported to give good results in bond dissociation energy calculations46,47,68–72 and the double-hybrid functional B2G-PLYP functional which has also been shown to perform well for thermochemical calculations.48,52,73 In all cases Grimme's D3 dispersion energy correction was employed, which has been shown to improve BDE (and thermochemical) values for most functionals50,52,74 and the aug-cc-pVTZ basis set was used.49 The reference values for ΔH and ΔG for the dissociation of the C–Br bond were taken from the previous work of Coote et al.65 The results are summarised in Table 2.
Cmpd | R–Br → R˙ + Br˙ | |||||
---|---|---|---|---|---|---|
B3-LYPa | BMKb | M06-2Xb | B2G-PLYPb | Litc | ||
a 6-31+G(d). b aug-cc-pVTZ-D3. c Calculated at the G3(MP2)-RAD level of theory at 298K in the gas phase; taken from ref. 65. | ||||||
MBriB | ΔG | 185.0 | 217.8 | 206.5 | 207.7 | 221.2 |
ΔH | 230.1 | 263.0 | 251.6 | 252.8 | 258.5 | |
MBriBA | ΔG | 208.9 | 239.8 | 226.4 | 228.7 | NA |
ΔH | 251.0 | 282.0 | 268.6 | 270.8 | NA | |
ΔGG | −23.9 | −22.0 | −19.9 | −21.0 | ||
MBriBA2 | ΔG | 195.4 | 228.3 | 216.1 | 216.2 | NA |
ΔH | 238.8 | 271.8 | 259.6 | 259.7 | NA | |
ΔGG | −10.4 | −10.5 | −9.6 | −8.5 | ||
MCliB | ΔG | 225.9 | 276.0 | 265.6 | 254.2 | 278.5 |
ΔH | 271.7 | 321.7 | 311.3 | 299.9 | 315.3 | |
MCliBA | ΔG | 254.0 | 301.2 | 289.9 | 279.6 | NA |
ΔH | 296.8 | 343.9 | 332.6 | 322.3 | NA |
The closest ΔG and ΔH values to those in the literature were obtained for the BMK functional (UHF) confirming that this functional is the low cost method of choice for BDE calculations. Irrespective of the absolute values, of particular interest are the relative values of the bond dissociation free energies between the ester and amide initiators. Remarkably consistent values were obtained (excluding B3LYP) with an average value for the BDFE of −21.7 kJ mol−1 with a mean absolute deviation of 1.25 kJ mol−1. BDFE values for ATRP initiators have been demonstrated to be the major determinant for the equilibrium constants for activation of the initiators by the Cu catalysts. All else being equal (under identical reaction conditions with identical reagents) the relative BDFE values can be used to gauge the relative reactivity of the initiator species. In this case a −21.7 kJ mol−1 free energy difference would correspond to the ester being approximately 6335 times (KATRP(MBriPA)/KATRP(MBriP) = 0.000158) more active than the amide. In contrast MBriPA2 with two methyl groups on the amide nitrogen, gave an average −9.53 kJ mol−1 free energy difference would correspond to the ester only being approximately 47 times (KATRP(MBriPA2)/KATRP(MBriP) = 0.0214) more active than the amide. It has been previously reported that the experimental KATRP values for the secondary ATRP initiators ethylbromopropanoate and 2-bromo-N,N-diethylpropanamidewere 0.30 and 0.044 respectively giving only a ≈7 fold difference in activity.75 Thus for MBriP (and by default EBriB) the relative differences in reactivity mean that the polymerization should not proceed in a controlled manner if at all; such a low initiation activity, coupled with a methacrylate monomer, would lead to very slow polymerisation and very poor initiator efficiency. The actual lower activity of the amides is not surprising due to the amides generally having lower radical stabilisation energies than the equivalent esters.64 What is surprising is the calculated relative magnitude of difference in ATRP activity between and MBriP and between MBriPA and MBripA2. One possible reason presents itself whilst considering the minimal energy conformations of the three molecules which are illustrated in Fig. 1. The MBriP and MBriPA2 C–Br bond angles to the plane of the CO2 and CON are circa 77° and 65° respectively; for MBriPA however this angle is circa 4°. This appears to be a result of the presence of a intramolecular H–Br hydrogen bond. Such bonds have been seen for many α-Br aromatic amides.76,77 This H-bond has the apparent effect of strengthening the C–Br and increasing its BDE; entropic effects do not appear to play a role.
![]() | ||
Fig. 1 Optimised minimum energy conformations (B3LYP/6-31+G(d)) with O![]() |
To determine the experimental relative KATRP values for the two initiators EBriB and PBriBA used in the polymerisation of OEGMA, we followed the method described by Matyjaszewski et al. whereby a function of the Cu(II) concentration (eqn (1)) was measured against time.78,79 The increase in Cu(II) arises from the persistent radical effect and the slope of the plot of F([Cu(II)]) versus time gives a value for KATRP according to eqn (2).
![]() | (1) |
![]() | (2) |
Thus degassed and sealed ethanol solutions of initiator:
Cu(I)Cl
:
bipyridine (1
:
1
:
2) were prepared, the degassed initiators (EBriB or PBriBA) were introduced by syringe and the absorption at 740 nm was monitored with time using a UV-vis spectrometer. The extinction coefficient taken as a reference to determine the concentration of the Cu(II) was that of the bipyridine Cu(II)Cl2 complex. Given that we used a mixed halide system in the ATRP process (a Br initiator with Cu(I)Cl) the assumption of a constant extinction coefficient for the Cu(II) complex may be invalid (since we used Cu(II)Cl2 as the baseline measurement for 100% Cu(II)) and thus the values obtained cannot be taken as absolute. However the relative values should be significant. Measurements gave KATRP of 8.32 × 10−6 for the EBriB and 5.37 × 10−7 for PBriBA indicating that the activity of the ester is 15.5 times greater than that of the amide (plots of F[Cu(II)] versus time can be found in ESI). This is a remarkably different relative reactivity than that calculated from DFT. We surmised he most likely reason for this disparity was differing solvent effects on the two initiators. To ascertain the effect of the solvent on the reaction we carried out further calculations at the BMK/aug-cc-pVTZ level using the SMD solvent model of Truhlar.54 Xylene was chosen as a non-polar solvent with minimal H-bond properties and ethanol as the reaction solvent. The results of the calculations of ΔG and ΔH are illustrated graphically in Fig. 2 and values for −ΔGG and relative KATRP values (K/K0) are given in Table 3. A significant drop in the enthalpies and free energies of the bond dissociation is observed on going from the gas phase to a xylene solvent for the MBriP and the MBriPA initiators whereas the MBriPA2 actually decreases in ΔH but increases slightly in ΔG.
![]() | ||
Fig. 2 Bond dissociation energies and free energies for model initiators in solvents calculated at the BMK/aug-cc-pVTZ//B3LYP/6-31+G(d) level of theory. |
Cmpd | Gas | Xylene | Ethanol | |
---|---|---|---|---|
a ΔΔGG = difference between ΔG for compound relative to ΔG for ester (MBriP). b K/K0 = ratio of KATRP for compound to KATRP for ester (MBriP). | ||||
MBriP | ΔΔGa | 0 | 0 | 0 |
K/K0b | 1 | 1 | 1 | |
MBriPA | ΔΔGa | −20.4 | −12.6 | −6.6 |
K/K0b | 0.000264 | 0.00626 | 0.0710 | |
MBriPA2 | ΔΔGa | −10.5 | −15.2 | −5.5 |
K/K0b | 0.0144 | 0.00216 | 0.107 |
For the ester however there is no significant difference in these values between xylene and ethanol. The amides though show further significant drops in ΔG and ΔH in ethanol. Thus the MBriPA and MBriPA2 initiators are circa 14 and 9 less reactive than the ester. The former value is in excellent agreement with the measured experimental relative reactivity for the PrBriPA initiator (K/K0 = 0.0646, circa 15.5 times less reactive than the ester).
Thus whilst solvation in a polar protic solvent dramatically increases the rate of activation of the amide structures for ATRP it is still less than that of the ester. Significant solvent effects have previously been observed (DMF vs. DMSO).59,80 Complexation of the Cu catalyst by the amide group(s) has also been put forward as an explanation for the low activity of amide initiators though experimental results are contradictory on this point55,58 and no conclusive evidence for complexation has been put forward to the best of our knowledge. To confirm that complexation of the amide with the copper complex was not significant we conducted UV-vis spectroscopic studies of the Cu(I)Cl, Cu(II)Cl2 and Cu(II)Br2 in the presence of an analogue of the amide initiator (N-methylpivalamide, Scheme 1) and noted no significant differences in the UV spectra (see ESI†) that might be indicative of amide binding to Cu. This supports Adams et al. conclusion that amide complexation is not significant;58 complexation of copper by amides and peptides is rare without deprotonation of the nitrogen or a free acid function or histidine being present.81–84
The ATRP of methacrylates from amide initiators should always lead to poor initiator efficiencies, broader than optimal polydispersities and inconstant radical concentrations as a consequence of the more rapid propagation of the ester relative to initiation of the amide.
This supports the experimental data of Adams et al. who detected unreacted amide initiators even at high conversions.58This difference in reactivity between the amide initiator and the propagating methacrylate species, can in itself explain most experimental results observed for amide initiators used in standard ATRP conditions to date. The difference in activity can be ameliorated to a certain extent by appropriate choice of solvent. This is not to say that termination of amide activity (such as observed by Habraken et al.)59 does not take place but we believe it to be a consequence of the slower overall reaction times for amide initiated reactions.
We further note that given the high BDE of the MCliBA (Table 2), if any deactivation of an amide initiator by Cu(II) chloride complexes takes place, then for all intents and purposes this deactivates the initiator to any significant reactions for the lifetime of the polymerisation.
![]() | ||
Fig. 3 13C NMR spectrum of PEI (Mn = 600) with assignments. Inset shows predicted spectrum for structure shown. |
![]() | ||
Fig. 4 1H NMR spectra and assignments of PEI (bottom) and PEI-Br (top). Assignments for PEI correspond to structure shown in Fig. 1. |
ID | [M]/[I] | M n,theo | SEC Mnb | SECTripleMnc | t (h) | Convd (%) | M n/Mwb |
---|---|---|---|---|---|---|---|
a M n,theo = [M]/[I] × Mn(0) × % conversion. b SEC Mn from RI response c against PMMA standards. c SEC Mn from triple-detection SEC. d Conversion from 1H NMR. | |||||||
P1 | 50 | — | 15![]() |
— | — | 1.23 | |
P2 | 50 | 13![]() |
27![]() |
26![]() |
83 | 1.40 | |
P3 | 50 | 10![]() |
47![]() |
— | 63 | 1.40 | |
P4 | 100 | 19![]() |
30![]() |
28![]() |
59 | 1.25 | |
P5 | 100 | 32![]() |
32![]() |
— | 99 | 1.27 |
![]() | ||
Fig. 6 (a) 1H NMR spectrum of PEI-graft-POEGMA (P1) sample. (b) SEC molecular weight distributions of PEI-Br, amide initiated POEGMA (Am1) and PEI-graft-POEGMA (P1). |
![]() | ||
Fig. 8 Particle size distribution of (a) iron oxide nanoparticles and (b) PEI-graft-POEGMA coated superparamagnetic iron oxide nanoparticles based on DLS measurements. |
As can be seen from Fig. 7, the particles had shown no specific shape but irregular morphologies no different from the unstabilised bare samples. High magnification imaging showed cubic, octahedral and spherical morphologies as is often obtained from chemical co-precipitation. X-ray diffraction analysis of the nanoparticles synthesized in the presence of PEI-graft-POEGMA, given in Fig. 9, shows the characteristic reflections of peaks assigned to magnetite and/or maghemite without any impurities of other iron oxide phases. High resolution TEM shows the lattice fringes with a d-spacing of 0.254 nm which can be indexed to 311 plane of magnetite, showing the crystalline nature of the product. Selected area diffraction pattern of the particles showed the existence of randomly oriented small crystals of the magnetite phase (ESI†). As was expected for the co-precipitation method there was no difference in the morphology or diffraction pattern between the particles prepared with and without PEI-graft-PEOGMA (ESI†). The magnetic properties of saturation magnetization, coercivity, and remanence of the PEI-graft-POEGMA coated and bare nano-particles were analyzed by vibrating sample magnetometry. Fig. 10 shows the corresponding hysteresis loop with no coercivity or remanence indicating that both samples have a superparamagnetic nature. The saturation magnetization of particles coated with PEI-graft-POEGMA was found to be 40.7 emu g−1 while for bare particles it is 48 emu g−1 which is less than bulk magnetite. The lowering of magnetization can be attributed to the presence of non-magnetic PEI-graft-POEGMA coating as the crystal structure, morphology and average size of the nanoparticles remain unchanged.92 Thermogravimetric analysis (TGA) (Fig. 11) showed 6% weight loss for uncoated superparamagnetic iron oxide nanoparticles due to the release of absorbed water from the nanoparticle surfaces. On the other hand PEI-graft-POEGMA coated iron oxide nanoparticles shows a further loss of 8.5% demonstrating the existence of a degradable organic component (the PEI-graft-POEGMA coating) on the magnetic iron oxide nanoparticles.
![]() | ||
Fig. 9 X-ray diffraction pattern for PEI-graft-POEGMA coated magnetite nanoparticles compared simulated pattern of magnetite. nanoparticles. |
The superparamagnetic magnetite nanoparticles prepared in the presence of PEI-graft-POEGMA (P1) were dispersed in distilled water by sonication after the synthesis. Although some settling occurred over time, the dispersed particles largely remained stable in suspension over days (ESI†). It should be noted that the bare particles without any surface modifications aggregated and precipitated in a matter of minutes.
Footnotes |
† Electronic supplementary information (ESI) available: Plots of Cu(II) function with time; UV-vis absorption spectra of Cu(I)Cl, Cu(II)Cl2 and Cu(II)Br2 solutions with DiMePiVA; X-ray diffraction pattern of PEI-graft-POEGMA stabilised nanoparticles; high magnification TEM images of magnetite nanoparticles; pictures of particle suspensions; X-ray diffraction pattern and TEM of bare (unstabilised) nanoparticles 2D 1H–13C and 1H–15N NMR spectra of PEI; inputs for DFT calculations; Cartesian co-ordinates of energy minimised conformations of initiators. See DOI: 10.1039/c3py01094e |
‡ These authors made equal contributions to the research described in this manuscript. |
This journal is © The Royal Society of Chemistry 2014 |