You are on page 1of 9

Articles

https://doi.org/10.1038/s41557-018-0159-8

Robust gold nanorods stabilized by bidentate


N-heterocyclic-carbene–thiolate ligands
Michelle J. MacLeod, Aaron J. Goodman, Hong-Zhou Ye   , Hung V.-T. Nguyen, Troy Van Voorhis and
Jeremiah A. Johnson   *

Although N-heterocyclic carbenes (NHCs) have demonstrated outstanding potential for use as surface anchors, synthetic chal-
lenges have limited their application to either large planar substrates or very small spherical nanoparticles. The development
of a strategy to graft NHCs onto non-spherical nanomaterials, such as gold nanorods, would greatly expand their utility as
surface ligands. Here, we use a bidentate thiolate–NHC–gold(i) complex that is easily grafted onto commercial cetyl trimeth-
ylammonium bromide-stabilized gold nanorods through ligand exchange. On mild reduction of the resulting surface-tethered
NHC–gold(i) complexes, the gold atom attached to the NHC complex is added to the surface as an adatom, thereby precluding
the need for reorganization of the underlying surface lattice upon NHC binding. The resulting thiolate–NHC-stabilized gold
nanorods are stable towards excess glutathione for up to six days, and under conditions with large variations in pH, high and
low temperatures, high salt concentrations, or in biological media and cell culture. We also demonstrate the utility of these
nanorods for in vitro photothermal therapy.

N
-heterocyclic carbenes (NHCs) have emerged as promis- installation of NHCs onto gold nanomaterials of any arbitrary size
ing ligands for surface chemistry1,2. NHC-stabilized palla- and shape is needed.
dium3–8, nickel9,10, platinum11, ruthenium12–16 and gold17–20 Here, we report an approach for the delivery of bidentate NHC–
nanoparticles have been studied in the context of catalysis, NHC- thiolate ligands to gold nanorod surfaces that breaks the above limi-
stabilized gold nanoparticles21,22 and planar surfaces23–25 have tations and yields robust nanorods (NHC@Au nanorods, Fig. 1b).
demonstrated potential for biomedical applications, and related Our strategy leverages the well-established ligand exchange of thio-
persistent carbene-stabilized Si surfaces have expanded the scope of lates onto CTAB-stabilized gold nanorods28 to bring masked NHCs
Si-functionalization methods26. Despite these successes, and unam- near the nanorod surface. Subsequent activation of the masked
biguous evidence that NHC monolayers can be more stable than NHC provides NHC@Au nanorods where the original nanorod
thiol-based monolayers24, there remain critical gaps in the chemistry size and shape are maintained, which suggests minimal nanorod
and scope of NHC-functionalized nanomaterials. For example, non- restructuring from the ligand addition34. We hypothesized that
spherical NHC-stabilized nanomaterials, such as gold nanorods, are installation of NHCs via reduction of NHC–gold(i) complexes teth-
conspicuously absent in the literature. The installation of NHCs ered to the nanorod surface would yield stable interfaces without
onto such substrates represents a unique synthetic challenge that, the need for reorganization of the underlying gold lattice to achieve
if overcome, could significantly expand the utility of NHC-based optimal NHC surface binding. This approach was inspired by the
interfaces. For example, thermally stable NHC-functionalized gold often observed etching of gold surfaces by free NHCs24,29,35 as well
nanorods could be applied in the context of photothermal therapy as work from Glorius, Fuchs and co-workers, which suggested that
(PTT), wherein laser irradiation of the nanorod induces local heat- when NHCs are deposited onto planar gold substrates under ultra-
ing and cell death27. high vacuum they tend to abstract a gold atom from the surface
Protocols for the synthesis of gold nanomaterials with precise lattice to generate translationally mobile NHC@Au adatom com-
sizes and shapes have been finely tuned over decades28. These plexes36. In addition, it may also slow the formation of bis-NHC
methods often involve the use of intermediate stabilizing ligands gold complexes, which have been observed on planar gold sub-
such as cetyl trimethylammonium bromide (CTAB) or citrate in strates37. Indeed, here we show that this ‘adatom addition’ method
water. Due to the limited stability of CTAB- and citrate-passiv- (Fig. 1b) leads to NHC@Au nanorods that are stable under a wide
ated nanomaterials towards changes in pH and electrolyte con- range of harsh conditions, including those where traditional NHC
centration, a subsequent ligand exchange step, usually with a and thiol@Au nanorods are unstable. Finally, in vitro cell culture
thiol-based ligand, is used before application of the nanorod28. In assays suggest that NHC@Au nanorods are promising materials for
the context of NHC surface chemistry, it would be ideal to sim- laser-induced PTT.
ply exchange CTAB or citrate from commercial nanorods with
an NHC of interest; however, existing ligand exchange protocols Results and discussion
(Fig. 1a) that involve the formation of free NHCs are unlikely to Masked bidentate NHC–thiolate design and synthesis. We began
prove successful with ligands such as CTAB or citrate in aque- our studies with the synthesis of poly(ethylene glycol) (PEG)-
ous solutions6–8,16,19,29,30. Moreover, direct reduction methods for based imidazolium salt 1-Br​ (Fig.  2a) via alkylation of 5-ethynyl-
the synthesis of NHC-stabilized nanoparticles have so far only 1-methyl-1H-imidazole (1​ ) with 2-bromoethyl 2-nitrobenzyl
provided relatively small (≤​15 nm) exclusively spherical nanoma- sulfide (2​) followed by copper-catalysed azide-alkyne cycloaddition
terials (Fig.  1a)11,12,17,18,21,22,31–33. Thus, a different strategy for the with azido-terminated 2 kDa PEG (3​) (58% yield over two steps;

Department of Chemistry, Massachusetts Institute of Technology, Cambridge, MA, USA. *e-mail: jaj2109@mit.edu

Nature Chemistry | www.nature.com/naturechemistry


Articles NATure CHemisTry

a R R thiolate component of 1-Br​, 1-HCO3​ and 1-Au​. In the presence of


R R R R R2 R2
excess CTAB, the nanorods were stable for at least 5 h under 365 nm
R
N N
R light (8 W desktop lamp), as determined by the lack of significant
N N N N S changes in the longitudinal surface plasmon resonance (SPR) band in
R R R R
Reduction Ligand exchange the UV–vis spectrum (Supplementary Fig. 10). When excess CTAB
Au was removed via centrifugation before irradiation, the resulting
X ‘CTAB-poor’ gold nanorods were unstable under the same irradia-
Ligand-stabilized tion conditions. However, exposure of ‘CTAB-poor’ gold nanorods
NHC-Au(I) NHC-stabilized gold nanoparticle
complex gold nanoparticle (here, with a to 1-Au​(Supplementary Fig. 11; based on ligand optimization stud-
thioether ligand) ies (Supplementary Fig.  12), 0.1 equiv. of masked bidentate ligand
b R3 to CTAB was used throughout this work) for 5 min before irradia-
N
NO2 R3
tion yielded stable nanorods. To validate the importance of thio-
CTAB@Au
nanorods, Br
Au N late photogeneration, we compared the stability of the commercial
R3
S hν N N CTAB@Au exposed to 1-Br​both with and without ultraviolet irra-
N N CTAB Reduction diation to the same CTAB@Au exposed to a previously reported21
exchange S S PEGylated imidazolium salt that lacks a thioether/thiolate com-
Au ponent. After dialysing to remove unbound ligand, only 1-Br​ with
Br ultraviolet irradiation yielded stable, soluble nanorods (Fig. 3a, +​UV,
1-Au 1-Au@Au NHC@Au-I +​S). Aurophilic interactions could potentially stabilize the nanorods
exposed to 1-Au​without the thiolate component49. As shown in
Fig. 1 | Strategies for installation of NHCs onto gold nanomaterials. a, Supplementary Fig. 13, the longitudinal SPR was significantly blue-
Previous methods involve the reduction of preformed or in situ generated shifted for CTAB@Au treated with 1-Au​and not irradiated with
(the latter from imidazolium tetrahaloaurate salts) NHC–metal (for ultraviolet light, while it was much less shifted when 1-Au​and ultra-
example, gold) complexes, or the displacement of weakly bound (for violet light were used together. These observations and transmission
example, thioether or amine) ligands with free NHCs. So far, it has not yet electron microscopy (TEM) measurements (Supplementary Fig. 13
been possible to install NHCs onto gold nanomaterials of arbitrary size and and Supplementary Table 1) suggest that 1-Au​on its own induces a
shape with these approaches. X, a coordinating ligand, typically a halogen change in the nanorod size and/or aspect ratio, while in situ photo-
such as Cl or Br; R and R2, alkyl, aryl or other substituents. b, Here, we generation of the thiolate component of 1-Au​inhibits these unde-
introduce a bidentate thiolate masked NHC strategy whereby exchange of sired alterations.
CTAB ligands on commercial CTAB@Au nanorods with a photogenerated Direct comparisons of the UV–vis spectra for CTAB@Au before
thiolate is followed by NHC installation. Shown is the synthesis of thiolate and after installation of the thiolates of 1-Br​ and 1-Au​ (nanorods
monolayer 1-Au@Au from photodeprotection of 1-Au​and its subsequent are named 1-Br@Au and 1-Au@Au, respectively) are provided
reduction to generate NHC@Au-I, which introduces a new surface-bound in Fig.  3b (see Supplementary Fig.  14 for 1-HCO3@Au). A rep-
NHC–gold adatom complex. The resulting bidentate thiolate–NHC- resentative TEM image of 1-Au@Au is provided in Fig.  3c (see
stabilized gold nanorods have the same size and shape as their parent Supplementary Table 1 for TEM-measured sizes and aspect ratios;
commercial nanorods and are robust towards a range of stringent see Supplementary Fig.  14 for TEM images of 1-Br@Au and
conditions. R3, triazole-conjugated polyethylene glycol. 1-HCO3@Au). In general, installation of our photogenerated thio-
late ligands produced slightly smaller nanorods compared to CTAB@
Au, as assessed by UV–vis and TEM (Supplementary Table 1). The
same phenomenon was also observed for a traditional PEG-thiol
see Supplementary Information for details; Supplementary Figs. 1– (PEG-SH) ligand (Supplementary Table  1 and Supplementary
4). Conversion of 1-Br​to gold(i) complex 1-Au​was achieved in Fig. 14). Fourier-transform infrared spectroscopy provides support
85% yield following methods reported by Nolan and co-workers for the presence of our PEG-based ligands (Supplementary Fig. 15);
for related imidazolium salts38. We also prepared the hydrogen however, it should be noted that we expect a small amount of CTAB
carbonate salt 1-HCO3​(see Supplementary Information, 90% to remain bound to the nanorods following ligand exchange in each
yield from 1-Br​) to compare how the NHC installation method case, which is commonly observed in ligand exchange protocols
impacts nanorod stability39. 1H NMR, 13C NMR and matrix-assisted using thiols50,51. Although ligand exchange onto ‘CTAB-poor’ gold
laser desorption ionization time-of-flight (MALDI–TOF) mass nanorods was possible (vide supra), the process of CTAB removal
spectrometry (MS) supported the structures of these compounds produced nanoparticle impurities (Supplementary Fig. 16), which
(Fig. 2b for 1-Au​; Supplementary Figs. 5–9). These bidentate ligands led us to abandon this approach in this work. In summary, the above
were designed to enable in situ photogeneration of the thiolate and results confirm that in situ photogeneration of thiolates introduces
subsequent ligand exchange onto commercially available CTAB- our ligands onto commercial CTAB@Au in water in a manner that
stabilized gold nanorods followed by NHC installation (Fig.  1b). minimally perturbs the nanorod size and aspect ratio.
Although NHC–thiolate bidentate ligands have not been used Next, we investigated methods for the installation of the NHC
in the context of surface functionalization, we reasoned that they components of our ligands onto these nanorod surfaces. To realize
would provide enhanced surface stability, as is known for related our adatom addition strategy, 1-Au@Au was treated with borane
multidentate ligands8,40–44. The ortho-nitrobenzyl protecting group tert-butylamine complex (2 equiv. to 1-Au​ ) in THF to provide
enables on-demand thiolate generation following irradiation with NHC@Au-I. For comparison to traditional methods for NHC
ultraviolet light45, while PEG is frequently used to impart aqueous installation, 1-Br@Au was exposed to potassium bis(trimethylsilyl)
solubility and biocompatibility to gold nanomaterials46–48. amide (KHMDS) in THF to provide NHC@Au-II, and 1-HCO3@
Au was stirred for 3 days in THF to provide NHC@Au-III. The
NHC@Au nanorod synthesis and characterization. Commercially UV–vis spectra for 1-Au@Au and NHC@Au-I were similar
available CTAB-stabilized (3  mM) 40  ×​ 10 nm gold nanorods (Fig.  4a) while the longitudinal SPR bands for NHC@Au-II and
(CTAB@Au) were used to optimize our ligand installation condi- NHC@Au-III (Supplementary Fig.  17) were slightly red- and
tions. We first confirmed the stability of CTAB@Au towards long- blueshifted compared to 1-Br@Au and 1-HCO3@Au, respectively,
wavelength ultraviolet light, which is required for unmasking of the which suggests that exposure to strong base (KHMDS) and/or the

Nature Chemistry | www.nature.com/naturechemistry


NATure CHemisTry Articles
a O 44 O 44

NO2
N N
(i) Br N N
S O2N O2N
N N
2 Br Au(SMe3)Cl
N N S N N S
N N K2CO3
O
(ii) N3 Au
44
1 3 1-Br Br 1-Au

b m c
O
O o
n 43
l [M – Br]+ = obs. 2,449.9
k m, n calc. 2,449.2
N
N j o
O2N
N
i d e
c
N N S f
a b h g 44 (C2H4O) }
Au
a
Br
i + CHCl3 CH2Cl2

d
j
k b l c
e gh f

8.4 8.0 7.6 7.2 6.8 6.4 6.0 5.6 5.2 4.8 4.4 4.0 3.6 3.2 2.8 1,500 2,000 2,500 3,000 3,500
Chemical shift (ppm) m/z

Fig. 2 | Synthesis and molecular characterization of PEGylated masked bidentate ligands. a, Synthesis of 1-Au​. Alkylation of commercially available 1​
followed by copper-catalysed azide-alkyne cycloaddition (CuAAC) to install azido-polyethylene glycol 3​yields imidazolium salt 1-Br​. Exposure of 1-Br​
to Au(SMe2)Cl in the presence of K2CO3 produces the macromolecular gold complex 1-Au​. In this work, 1-Br​is used as the ‘traditional’ NHC source for
subsequent surface studies (that is, the free NHC is generated by deprotonation of 1-Br​). In contrast, 1-Au​is used for adatom addition. b,c, 1H NMR
(CDCl3, 500 MHz) spectrum (b) and MALDI–TOF spectrum (c) for 1-Au​. The data corroborate the proposed structure of 1-Au​.

a b c
CTAB@Au
+UV, +S
1-Au@Au
–UV, +S
1-Br@Au
+UV, –S
Absorbance (a.u.)
Absorbance (a.u.)

500 600 700 800 900 1,000 500 600 700 800 900 1,000
Wavelength (nm) Wavelength (nm)

Fig. 3 | Installation of masked NHC–thiolate ligands onto commercial CTAB@Au nanorods. a, UV–vis spectra for CTAB@Au exposed to 1-Br​before (–UV,
+​S) and after (+​UV, +​S) exposure to ultraviolet light. The results demonstrate that photogeneration of the thiolate component of 1-Br​ (+​UV, +​S) is critical
for generating stable gold nanorods (that is, the thioether of 1-Br​doesn’t stabilize the nanorods on its own). Also shown are data collected for a non-thiol-
containing PEGylated imidazolium salt where both N substituents are methyl (+​UV, –S), affirming the importance of the thiolate component. b, UV–vis
spectra of CTAB@Au, 1-Br​ and 1-Au​thiolate-exchanged nanorods. Thiolate exchange leads to minimal changes in the transverse and longitudinal SPR
bands compared to the UV–vis spectra of the parent CTAB@Au, suggesting that this process does not dramatically impact the nanorod size and shape. c,
TEM image of 1-Au@Au. Scale bar, 30 nm. The image confirms the presence of uniform nanorods following ligand exchange.

free NHC induced some reorganization of the nanorod structure. Stability and density functional theory studies reveal the impact
Nonetheless, TEM indicated that the NHC installation method had of adatom addition. To assess the stability of bidentate NHC–thi-
little obvious effect on the size and aspect ratio of the nanorods olate-stabilized nanorods NHC@Au-I-III, we exposed each sample
(Fig. 4b, Supplementary Fig. 17 and Supplementary Table 1). and its precursor (1-Au@Au, 1-Br@Au, 1-HCO3@Au) to a wide

Nature Chemistry | www.nature.com/naturechemistry


Articles NATure CHemisTry

a As an extreme test of the ability of our ligands to protect the


R3
gold nanorod surface, we exposed each sample to aqueous solutions
N
R3
of KCN at various concentrations (Supplementary Figs.  33–35).
Br
Au N Under these conditions, only NHC@Au-I and NHC@Au-II dis-
TBAB
N N played some resistance to etching in 50 mM KCN for 10 min. These
THF observations suggest that although aurophilic interactions may sta-
S S
bilize 1-Au@Au, there is additional stability gained by the bidentate
NHC–thiolate ligand structure. In addition, NHC@Au-I appears to
perform the best under the most stringent conditions, suggesting
1-Au@Au NHC@Au-I that the addition of new gold atoms to the surface is advantageous.
To test whether or not increasing the ligand density would
b
1-Au@Au enhance the etching resistance of NHC@Au-I, 1-Au@Au was
treated with an additional 0.034 or 0.085 equiv. (compared to CTAB
NHC@Au-I
in CTAB@Au nanorods) of 1-Au​and exposed to ultraviolet light.
The longitudinal SPR bands for both samples were slightly blue-
shifted compared to the parent 1-Au@Au (Supplementary Fig. 36).
Absorbance (a.u.)

Reduction followed by TEM imaging (Supplementary Figs.  37


and 38) revealed that the nanorod shape was not significantly
modified by exposure to additional 1-Au​. However, the stability
of these nanorods towards KCN etching was markedly improved:
both samples were completely resistant to 50 mM KCN for 10 min
with significant, though greatly redshifted, SPR bands after 25 min
(Supplementary Figs.  39 and 40, respectively). Once again high-
lighting the importance of adding a gold atom to the surface, the
500 600 700 800 900 1,000 non-reduced samples were rapidly etched under these condi-
Wavelength (nm) tions regardless of the presence of additional thiolate ligand 1-Au​
(Supplementary Figs. 41 and 42, respectively).
c
Density functional theory (DFT) calculations were used to eval-
uate the binding energetics of analogues of our bidentate ligands to
a fixed 19-gold-atom cluster where the NHC-bound gold atom was
allowed to relax. Docking of the thiyl component of model ligand
1b+ (Fig. 5a) to the model gold surface produced 1b+@Au with a
binding energy of 1.78 eV, which reflects the Au–S bond strength.
Docking of the tethered NHC 1b@Au (Fig. 5b) provided NHCb@
Au-II with an additional binding energy of 1.57 eV. Notably, the lat-
ter value, which reflects the NHC–gold bond strength, is relatively
small, perhaps due to steric strain at the interface. The overall bind-
ing energy for this bidentate ligand that lacks adatom addition capa-
bility was 3.35 eV.
Docking of 1b-Au-Br (Fig. 5c), an analogue of 1-Au​, provided
1b-Au-Br@Au with a binding energy of 1.49 eV. This value is some-
what less than the value for 1b@Au; however, docking of the reduced
Fig. 4 | Characterization of NHC@Au-I. a, Synthesis of NHC@Au-I from ligand, 1b-Au@Au (Fig. 5d), led to NHCb@Au-I with an additional
1-Au@Au via reduction of the surface-tethered NHC complex to achieve binding energy of 2.45 eV. This value is within the range of bind-
adatom addition. TBAB, tert-butylamine borane; THF, tetrahydrofuran; ing energies calculated for other NHC–gold adatom complexes
R3, triazole-conjugated polyethylene glycol. b, UV–vis spectra for 1-Au@ on gold substrates (from ~1.7 eV to 2.69 eV)52. Thus, the overall
Au and NHC@Au-I showing that the size and shape of the nanorods are bidentate ligand binding energy in NHC@Au-I (3.94 eV) is 0.59 eV
maintained on conversion of 1-Au@Au to NHC@Au-I (that is, adatom greater than in NHC@Au-II (3.35 eV). Moreover, the NHC–gold
addition does not dramatically affect the nanorod optical properties). c, bond was strengthened significantly to 2.84 eV in NHCb@Au-I
TEM image of NHC@Au-I confirming the presence of uniform nanorods (Supplementary Fig.  43), compared to 1.57 eV in NHCb@Au-II,
following adatom addition (for comparison to 1-Au@Au see Fig. 3c). Scale which suggests that the addition of an adatom in the former leads
bar, 30 nm. to less steric strain. Notably, the NHC–gold bond in NHCb@Au-I
(2.84 eV) is significantly stronger than the bond between the NHC–
gold complex and the underlying gold surface (2.45 eV); desorption
range of pH values (from 2 to 14), high electrolyte concentrations of the ligand should involve removal of the newly added adatom.
(up to 1 M NaCl) and gold etching conditions (25, 50 and 100 mM Glorius, Fuchs and co-workers proposed that on planar gold
KCN) (Supplementary Figs.  18–35)48,50,51. In these studies, both substrates, NHC-mediated extraction of gold atoms from the sur-
NHC@Au-I and its precursor 1-Au@Au displayed superior stabil- face to generate new adatoms involved concerted movement of an
ity to all of the other samples. Minimal changes in SPR bands for underlying lattice atom to the surface36. In such systems with a large
NHC@Au-I and 1-Au@Au were observed after three days (longest ratio of gold atoms to NHCs, adatom addition may not lead to sub-
examination time) in aqueous solution at pH 2 and 14 and after up stantial increases in binding energy; however, extraction of surface
to 16 days in aqueous 1 M NaCl (Supplementary Figs. 18–20), while gold atoms by NHCs on NHC@Au nanorods may leave vacancies
NHC@Au-II and NHC@Au-III as well as their precursors 1-Br@ within the bulk nanorod. Previous work has shown that there is a
Au and 1-HCO3@Au, showed long-term stability (minimal change ~0.7–1 eV energy penalty for removal of a bulk gold atom, with sub-
in the SPR band over several days) within a narrower pH range of 7 stantial dependence on geometry53,54. To calibrate our simulations,
to 11 and in ≤​100 mM aqueous NaCl (Supplementary Figs. 21–32). we calculated that the formation of a such a vacancy in a cluster of

Nature Chemistry | www.nature.com/naturechemistry


NATure CHemisTry Articles
a b

19 Au

–1.78 eV –1.57 eV

1b+ Au atom
extracted
+
1b @Au 1b@Au NHCb@Au-II

c d

19 Au

–1.49 eV –2.45 eV

1b-Au-Br
New Au
atom
1b-Au@Au 1b-Au(0)@Au NHCb@Au-I

Fig. 5 | DFT calculations. a, A non-PEGylated thiyl-imidazolium salt (1b+) binds to a model 19-atom gold substrate with a binding energy of 1.78 eV to
form 1b+@Au. Note that the bromide counterion is neglected for simplicity. b, Deprotonation of 1b+@Au provides the surface-tethered free NHC 1b@Au,
which then binds to the substrate with an additional binding energy of 1.57 eV to yield NHCb@Au-II, which is analogous to NHC@Au-II. A gold atom from
the underlying lattice is extracted to accommodate the NHC (red arrow). The overall binding energy of the bidentate ligand is calculated to be 3.35 eV.
c, A non-PEGylated thiyl-NHC–Au–Br complex (1b-Au-Br) binds to a model 19-atom gold substrate with a binding energy of 1.49 eV to form 1b-Au-Br@
Au. Reduction of 1b-Au@Au provides the surface-tethered NHC–gold(0) complex, which then binds to the substrate with an additional binding energy
of 2.45 eV to yield NHCb@Au-I, which is analogous to NHC@Au-I. Addition of a new gold atom (green arrow) precludes the need for modification of the
underlying lattice. The overall binding energy of the bidentate ligand is calculated to be 3.94 eV.

19 gold atoms would incur a 1.89 eV energy penalty (Supplementary light into heat. In addition, NHC@Au-I was stable for at least three
Fig.  43). Thus, we propose that the enhanced stability of NHC@ cycles of irradiation (Fig. 6e). Next, MCF7 human breast adenocar-
Au-I and NHCb@Au-I compared to the traditional NHC-based cinoma cell viability assays were conducted to assess the potential of
ligands is due at least in part to the installation of one adatom per NHC@Au-I for in vitro PTT. Cells were irradiated under the same
NHC ligand, which precludes the need for lattice reorganization. To conditions used above (830 nm laser, 6 W cm−2 for 8 min) to con-
our knowledge, this adatom addition approach represents a funda- firm that laser irradiation alone did not induce cell death (Fig. 6f).
mentally different way to install NHC ligands onto surfaces. Incubation of NHC@Au-I in cell culture media for 3 days at 37 °C
led to no change in the longitudinal SPR band (the transverse band
Stability under bio-related conditions and in vitro PTT studies overlaps with cell media), which suggests that the nanorods are
using NHC@Au-I. Given the outstanding robustness of NHC@ stable in cell media (Supplementary Fig. 44)19. Cell viability assays
Au-I, we sought to explore its preliminary biological applications. (CellTiter-Glo) indicated that doses of 3.8 and 7.5 μ​g of NHC@Au-I
Both thiol- and NHC-stabilized gold nanomaterials are susceptible in the absence of laser irradiation were well tolerated, with 91 ±​  16%
to degradation by biological thiols such as glutathione (GSH); to our (P = 0.59) and 90 ±​  10% (P = 0.46) cell viability, respectively (Fig. 6f
knowledge, there are no examples of NHC-stabilized gold nanoma- and Supplementary Fig.  45). Laser irradiation of cells exposed
terials that have shown long-term stability towards 2 mM GSH for to NHC@Au-I led to significant cell killing: cell viabilities were
24 h (in vivo, GSH concentration ranges from μ​M to ~10 mM)21,22. 24 ±​  30% (P =​ 0.026) for the 3.8 μ​g dose and 6 ±​  1% (P =​  0.0001) for
Remarkably, when NHC@Au-I was exposed to 4 mM GSH in water the 7.5 μ​g dose (Fig. 6f). Thus, the combination of NHC@Au-I and
there was almost no change in the SPR bands over 6 days (Fig. 6a), laser irradiation leads to selective cell killing. These results show
which suggests that these nanorods are stable towards GSH. that NHC-functionalized interfaces can be applied in cellular deliv-
As discussed above, for use in PTT, nanorods must be stable at ery assays; they also suggest that NHC@Au-I could be a viable can-
high temperatures55–57. On the other hand, very low temperatures didate for translation to in vivo PTT.
are often encountered in biological applications (for example, long-
term cell storage). To test the temperature stability of our nanorods, Conclusions
aqueous solutions of NHC@Au-I were cooled to −​78 °C or heated In summary, we have demonstrated a bidentate ligand strategy for
to 95 °C for 5 h. Under these conditions, no significant changes the generation of NHC@Au surface linkages that overcomes the size,
in the SPR bands were observed (Fig.  6b). In contrast, analogous shape and solvent limitations of previous methods. Three methods
PEG-SH-stabilized nanorods (PEG-S@Au) experienced extensive for the installation of NHCs via this strategy were experimentally
degradation at both low and high temperatures (Fig. 6c). It should tested; and an ‘adatom addition’ approach involving reduction58 of
be noted that the temperature stability of PEG-S@Au could poten- thiolate-tethered NHC–Au complexes proved to be the most suc-
tially be augmented by increasing the amount of PEG-SH used in cessful, providing robust gold nanorods that were stable towards
the CTAB exchange step44. Nevertheless, our data show that at the wide pH ranges, high salt concentration, gold etching conditions,
same ligand concentration, NHC@Au-I is significantly more stable GSH, and high and low temperatures. Moreover, these materials
towards temperature variations than PEG-S@Au. were effective for in vitro PTT. We attribute the unprecedented sta-
Progressing to preliminary PTT experiments, we irradiated solu- bility of these NHC-functionalized surfaces to the adatom addition
tions of NHC@Au-I with a near-infrared (830 nm) laser using a flu- installation method, which avoids the need for reorganization of the
ence of 6 W cm−2 for 8 min. As expected, the solution temperature nanorod surface to enable optimal NHC binding. This work signifi-
increased linearly with NHC@Au-I concentration (Fig. 6d), which cantly expands both the synthesis and applications of NHC surface
confirms that our nanorods are capable of converting absorbed anchors. Moreover, we believe that the concept of adatom addition

Nature Chemistry | www.nature.com/naturechemistry


Articles NATure CHemisTry

a b c
Without GSH RT, 5 h RT, 5 h
+GSH, 0 h –78 °C, 5 h –78 °C, 5 h
Absorbance (a.u.) +GSH, 8 h 95 °C, 5 h 95 °C, 5 h

Absorbance (a.u.)

Absorbance (a.u.)
+GSH, 1 d
+GSH, 6 d

500 600 700 800 900 1,000 500 600 700 800 900 1,000 500 600 700 800 900 1,000
Wavelength (nm) Wavelength (nm) Wavelength (nm)

d 30 e Concentration (μg ml–1): 100 f


NS
NS
100
Δ Temperature (°C)

20 Absorbance (a.u.)

Cell viability (%)


80
50
60 *
10 40
25
20
***
0 0
0 25 50 100 500 600 700 800 900 1,000 Dose (μg): 0 3.8 3.8 7.5 7.5
Concentration (μg ml–1) +hν –hν +hν –hν +hν
Wavelength (nm)

Fig. 6 | Preliminary biological investigations using NHC@Au-I. a, UV–vis spectra for NHC@Au-I dissolved in 4 mM GSH, displaying the stability ofthe
nanorods for up to 6 days (longest time tested). b, UV–vis spectra for NHC@Au-I displaying the stability of the nanorods in water at high and low
temperatures. RT, room temperature. c, UV–vis spectra for PEG-S@Au showing the nanorods’ lack of thermal stability under the same conditions as
b. d, Change in water temperature after irradiation of NHC@Au-I with a near-infrared (830 nm) laser using a fluence of 6 W cm−2 for 8 min. Error bars
correspond to s.d. of three independent measurements. e, UV–vis spectra for NHC@Au-I at various concentrations before and after irradiation for three
cycles; no changes in the spectra are observed with each irradiation cycle, which suggests that the nanorods are stable in these conditions. f, MCF7 in
vitro cell viability studies as determined by CellTiter-Glo. Laser irradiation (dose of 0 μ​g ml−1) or exposure to 100 μ​l of 38 μ​g ml−1 or 75 μ​g ml−1 of NHC@Au-I
had no impact on cell viability. In contrast, laser irradiation of cells exposed to 100 μ​l of 38 μ​g ml−1 or 75 μ​g ml−1 NHC@Au-I induced significant cell killing.
Statistical significance was assessed by a two-tailed t-test (n = 3 independent samples). NS, not significant, *P <​ 0.05, ***P <​ 0.001.

6. Richter, C., Schaepe, K., Glorius, F. & Ravoo, B. J. Tailor-made N-heterocyclic


to install NHC ligands onto metallic surfaces may provide a general carbenes for nanoparticle stabilization. Chem. Commun. 50, 3204–3207 (2014).
route to highly stabilized interfaces. 7. Ferry, A. et al. Negatively charged N-heterocyclic carbene-stabilized Pd
and Au nanoparticles and efficient catalysis in water. ACS Catal. 5,
Reporting Summary. Further information on experimental design 5414–5420 (2015).
is available in the Nature Research Reporting Summary linked to 8. Rühling, A. et al. Modular bidentate hybrid NHC–thioether ligands for the
this article. stabilization of palladium nanoparticles in various solvents. Angew. Chem. Int.
Ed. 55, 5856–5860 (2016).
9. Soulé, J.-F., Miyamura, H. & Kobayashi, S. Copolymer-incarcerated nickel
Data availability nanoparticles with N-heterocyclic carbene precursors as active cross-linking
All data supporting the findings of this study are available within the Article and
agents for Corriu–Kumada–Tamao reaction. J. Am. Chem. Soc. 135,
its Supplementary Information, and/or from the corresponding author upon
10602–10605 (2013).
reasonable request.
10. de los Bernardos, M. D., Pérez-Rodríguez, S., Gual, A., Claver, C. & Godard,
C. Facile synthesis of NHC-stabilized Ni nanoparticles and their catalytic
Received: 3 April 2018; Accepted: 14 September 2018; application in the Z-selective hydrogenation of alkynes. Chem. Commun. 53,
Published: xx xx xxxx 7894–7897 (2017).
11. Lara, P., Suárez, A., Collière, V., Philippot, K. & Chaudret, B. Platinum
References N-heterocyclic carbene nanoparticles as new and effective catalysts for the
1. Zhukhovitskiy, A. V., MacLeod, M. J. & Johnson, J. A. Carbene ligands in selective hydrogenation of nitroaromatics. ChemCatChem 6, 87–90 (2014).
surface chemistry: from stabilization of discrete elemental allotropes to 12. Lara, P. et al. Ruthenium nanoparticles stabilized by N-heterocyclic carbenes:
modification of nanoscale and bulk substrates. Chem. Rev. 115, 11503–11532 ligand location and influence on reactivity. Angew. Chem. Int. Ed. 50,
(2015). 12080–12084 (2011).
2. Hopkinson, M. N., Richter, C., Schedler, M. & Glorius, F. An overview of 13. Gonzalez-Galvez, D. et al. NHC-stabilized ruthenium nanoparticles as
N-heterocyclic carbenes. Nature 510, 485–496 (2014). new catalysts for the hydrogenation of aromatics. Catal. Sci. Tech. 3,
3. Ranganath, K. V. S., Kloesges, J., Schäfer, A. H. & Glorius, F. Asymmetric 99–105 (2013).
nanocatalysis: N-heterocyclic carbenes as chiral modifiers of Fe3O4/Pd 14. Lara, P. et al. NHC-stabilized Ru nanoparticles: synthesis and surface studies.
nanoparticles. Angew. Chem. Int. Ed. 49, 7786–7789 (2010). Nano-Structures Nano-Objects 6, 39–45 (2016).
4. Ranganath, K. V. S., Schäfer, A. H. & Glorius, F. Comparison of 15. Martínez-Prieto, L. M. et al. Long-chain NHC-stabilized RuNPs as versatile
superparamagnetic Fe3O4-supported N-heterocyclic carbene-based catalysts catalysts for one-pot oxidation/hydrogenation reactions. Chem. Commun. 52,
for enantioselective allylation. ChemCatChem 3, 1889–1891 (2011). 4768–4771 (2016).
5. Serpell, C. J., Cookson, J., Thompson, A. L., Brown, C. M. & Beer, P. D. 16. Ernst, J. B., Muratsugu, S., Wang, F., Tada, M. & Glorius, F. Tunable
Haloaurate and halopalladate imidazolium salts: structures, properties, and heterogeneous catalysis: N-heterocyclic carbenes as ligands for supported
use as precursors for catalytic metal nanoparticles. Dalton Trans. 42, heterogeneous Ru/K-Al2O3 catalysts to tune reactivity and selectivity. J. Am.
1385–1393 (2013). Chem. Soc. 138, 10718–10721 (2016).

Nature Chemistry | www.nature.com/naturechemistry


NATure CHemisTry Articles
17. Song, S. G. et al. N-Heterocyclic carbene-based conducting polymer–gold 42. Oh, E., Susumu, K., Goswami, R. & Mattoussi, H. One-phase synthesis of
nanoparticle hybrids and their catalytic application. Macromolecules 47, water-soluble gold nanoparticles with control over size and surface
6566–6571 (2014). functionalities. Langmuir 26, 7604–7613 (2010).
18. Crespo, J. et al. Ultrasmall NHC-coated gold nanoparticles obtained through 43. Man, R. W. Y. et al. Ultrastable gold nanoparticles modified by bidentate
solvent free thermolysis of organometallic Au(i) complexes. Dalton Trans. 43, N-heterocyclic carbene ligands. J. Am. Chem. Soc. 140, 1576–1579 (2018).
15713–15718 (2014). 44. Cao, Z. et al. Chelating N-heterocyclic carbene ligands enable tuning of
19. Cao, Z. et al. A molecular surface functionalization approach to tuning electrocatalytic CO2 reduction to formate and carbon monoxide: surface
nanoparticle electrocatalysts for carbon dioxide reduction. J. Am. Chem. Soc. organometallic chemistry. Angew. Chem., Int. Ed. 57, 4981–4985 (2018).
138, 8120–8125 (2016). 45. Pillar, V. N. R. Photo-removable protecting groups in organic-synthesis.
20. Ye, R. et al. Supported Au nanoparticles with N-heterocyclic carbene ligands Synthesis 1980, 1–26 1980).
as active and stable heterogeneous catalysts for lactonization. J. Am. Chem. 46. Dreaden, E. C., Alkilany, A. M., Huang, X., Murphy, C. J. & El-Sayed, M. A.
Soc. 140, 4144–4149 (2018). The golden age: gold nanoparticles for biomedicine. Chem. Soc. Rev. 41,
21. MacLeod, M. J. & Johnson, J. A. PEGylated N-heterocyclic carbene anchors 2740–2779 (2012).
designed to stabilize gold nanoparticles in biologically relevant media. J. Am. 47. Otsuka, H., Nagasaki, Y. & Kataoka, K. PEGylated nanoparticles for biological
Chem. Soc. 137, 7974–7977 (2015). and pharmaceutical applications. Adv. Drug Deliv. Rev. 64, 246–255 (2012).
22. Salorinne, K. et al. Water-soluble N-heterocyclic carbene-protected gold 48. Locatelli, E., Monaco, I. & Franchini, M. C. Surface modifications of gold
nanoparticles: size-controlled synthesis, stability, and optical properties. nanorods for applications in nanomedicine. RSC Adv. 5, 21681–21699 (2015).
Angew. Chem. Int. Ed. 56, 6198–6202 (2017). 49. Liu, H.-X., He, X. & Zhao, L. Metallamacrocycle-modified gold nanoparticles: a
23. Zhukhovitskiy, A. V., Mavros, M. G., Van Voorhis, T. & Johnson, J. A. new pathway for surface functionalization. Chem. Commun. 50, 971–974 (2014).
Addressable carbene anchors for gold surfaces. J. Am. Chem. Soc. 135, 50. Schulz, F. et al. Effective PEGylation of gold nanorods. Nanoscale 8,
7418–7421 (2013). 7296–7308 (2016).
24. Crudden, C. M. et al. Ultra stable self-assembled monolayers of 51. Khlebtsov, N. & Dykman, L. Biodistribution and toxicity of engineered gold
N-heterocyclic carbenes on gold. Nat. Chem. 6, 409–414 (2014). nanoparticles: a review of in vitro and in vivo studies. Chem. Soc. Rev. 40,
25. Crudden, C. M. et al. Simple direct formation of self-assembled 1647–1671 (2011).
N-heterocyclic carbene monolayers on gold and their application in 52. Tang, Q. & Jiang, D. E. Comprehensive view of the ligand gold interface from
biosensing. Nat. Commun. 7, 12654 (2016). first principles. Chem. Mater. 29, 6908–6915 (2017).
26. Zhukhovitskiy, A. V. et al. Reactions of persistent carbenes with hydrogen- 53. Schaefer, H. E. Investigation of thermal-equilibrium vacancies in metals by
terminated silicon surfaces. J. Am. Chem. Soc. 138, 8639–8652 (2016). positron-annihilation. Phys. Status Solidi A 102, 47–65 (1987).
27. Mackey, M. A., Ali, M. R. K., Austin, L. A., Near, R. D. & El-Sayed, M. A. 54. Qi, W. H. & Wang, M. P. Vacancy formation energy of small particles. J.
The most effective gold nanorod size for plasmonic photothermal therapy: Mater. Sci. 39, 2529–2530 (2004).
theory and in vitro experiments. J. Phys. Chem. B 118, 1319–1326 (2014). 55. Huang, X. H., El-Sayed, I. H., Qian, W. & El-Sayed, M. A. Cancer cell
28. Grzelczak, M., Pérez-Juste, J., Mulvaney, P. & Liz-Marzán, L. M. Shape control imaging and photothermal therapy in the near-infrared region by using gold
in gold nanoparticle synthesis. Chem. Soc. Rev. 37, 1783–1791 (2008). nanorods. J. Am. Chem. Soc. 128, 2115–2120 (2006).
29. Hurst, E. C., Wilson, K., Fairlamb, I. J. S. & Chechik, V. N-Heterocyclic 56. Huang, X. H. & El-Sayed, M. A. Gold nanoparticles: optical properties and
carbene coated metal nanoparticles. New J. Chem. 33, 1837–1840 (2009). implementations in cancer diagnosis and photothermal therapy. J. Adv. Res. 1,
30. Möller, N. et al. Stabilization of high oxidation state upconversion 13–28 (2010).
nanoparticles by N-heterocyclic carbenes. Angew. Chem. Int. Ed. 56, 57. Yang, X., Yang, M. X., Pang, B., Vara, M. & Xia, Y. N. Gold nanomaterials at
4356–4360 (2017). work in biomedicine. Chem. Rev. 115, 10410–10488 (2015).
31. Vignolle, J. & Tilley, T. D. N-Heterocyclic carbene-stabilized gold 58. Doud, E. A. et al. In situ formation of N-heterocyclic carbene-bound
nanoparticles and their assembly into 3D superlattices. Chem. Commun. single-molecule junctions. J. Am. Chem. Soc. 140, 8944–8949 (2018).
2009, 7230–7232 (2009).
32. Ling, X., Roland, S. & Pileni, M. P. Supracrystals of N-heterocyclic
carbene-coated Au nanocrystals. Chem. Mater. 27, 414–423 (2015). Acknowledgements
The authors thank the National Science Foundation (CHE-1351646) for support of this
33. Bridonneau, N. et al. N-Heterocyclic carbene-stabilized gold nanoparticles
work.
with tunable sizes. Dalton Trans. 47, 6850–6859 (2018).
34. Torrelles, X. et al. Solving the long-standing controversy of long-chain
alkanethiols surface structure on Au(111). J. Phys. Chem. C 122, Author contributions
3893–3902 (2018). M.J.M. and J.A.J. conceived the idea. M.J.M. conducted all synthesis and characterization
35. Rodríguez-Castillo, M. et al. Reactivity of gold nanoparticles towards studies. A.J.G. conducted laser irradiation experiments. H.Y. and T.V.V. conducted
N-heterocyclic carbenes. Dalton Trans. 43, 5978–5982 (2014). DFT calculations. H.V.-T.N. conducted cell culture assays. M.J.M. and J.A.J. wrote the
36. Wang, G. Q. et al. Ballbot-type motion of N-heterocyclic carbenes on gold manuscript. All authors read and revised the manuscript.
surfaces. Nat. Chem. 9, 152–156 (2017).
37. Jiang, L. et al. N-Heterocyclic carbenes on close-packed coinage metal
surfaces: bis-carbene metal adatom bonding scheme of monolayer films on Competing interests
Au, Ag and Cu. Chem. Sci. 8, 8301–8308 (2017). The authors declare no competing interests.
38. Collado, A., Gomez-Suarez, A., Martin, A. R., Slawin, A. M. Z. & Nolan, S. P.
Straightforward synthesis of [Au(NHC)X] (NHC =​  N-heterocyclic carbene,
X =​ Cl, Br, I) complexes. Chem. Commun. 49, 5541–5543 (2013). Additional information
39. Fèvre, M. et al. Imidazolium hydrogen carbonates versus imidazolium Supplementary information is available for this paper at https://doi.org/10.1038/
carboxylates as organic precatalysts for N-heterocyclic carbene catalyzed s41557-018-0159-8.
reactions. J. Org. Chem. 77, 10135–10144 (2012). Reprints and permissions information is available at www.nature.com/reprints.
40. Li, Z., Jin, R. C., Mirkin, C. A. & Letsinger, R. L. Multiple thiol-anchor capped Correspondence and requests for materials should be addressed to J.A.J.
DNA-gold nanoparticle conjugates. Nucleic Acids Res. 30, 1558–1562 (2002).
41. Oh, E. et al. Colloidal stability of gold nanoparticles coated with multithiol- Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in
poly(ethylene glycol) ligands: importance of structural constraints of the published maps and institutional affiliations.
sulfur anchoring groups. J. Phys. Chem. C 117, 18947–18956 (2013). © The Author(s), under exclusive licence to Springer Nature Limited 2018

Nature Chemistry | www.nature.com/naturechemistry


nature research | reporting summary
Corresponding author(s): Jeremiah A. Johnson

Reporting Summary
Nature Research wishes to improve the reproducibility of the work that we publish. This form provides structure for consistency and transparency
in reporting. For further information on Nature Research policies, see Authors & Referees and the Editorial Policy Checklist.

Statistical parameters
When statistical analyses are reported, confirm that the following items are present in the relevant location (e.g. figure legend, table legend, main
text, or Methods section).
n/a Confirmed
The exact sample size (n) for each experimental group/condition, given as a discrete number and unit of measurement
An indication of whether measurements were taken from distinct samples or whether the same sample was measured repeatedly
The statistical test(s) used AND whether they are one- or two-sided
Only common tests should be described solely by name; describe more complex techniques in the Methods section.

A description of all covariates tested


A description of any assumptions or corrections, such as tests of normality and adjustment for multiple comparisons
A full description of the statistics including central tendency (e.g. means) or other basic estimates (e.g. regression coefficient) AND
variation (e.g. standard deviation) or associated estimates of uncertainty (e.g. confidence intervals)

For null hypothesis testing, the test statistic (e.g. F, t, r) with confidence intervals, effect sizes, degrees of freedom and P value noted
Give P values as exact values whenever suitable.

For Bayesian analysis, information on the choice of priors and Markov chain Monte Carlo settings
For hierarchical and complex designs, identification of the appropriate level for tests and full reporting of outcomes
Estimates of effect sizes (e.g. Cohen's d, Pearson's r), indicating how they were calculated

Clearly defined error bars


State explicitly what error bars represent (e.g. SD, SE, CI)

Our web collection on statistics for biologists may be useful.

Software and code


Policy information about availability of computer code
Data collection for TEM: Microscope User Interface, Tecnai (version 4.5.1 Build 8403) & Gatan Digital Micrograph, Gatan Microscopy Suite (version
2.11.1408.0); for DFT: Q-Chem (version 5.1); for UV-vis measurements: Cary WinUV (version 5.0); for cell viability (plate reader): i-Control
1.10

Data analysis for TEM: ImageJ64 (1.48v); for data plotting: Microsoft Excel (version 16.16.1) and Matlab (version 9.3); for NMR: Mnova (version 11.0)
For manuscripts utilizing custom algorithms or software that are central to the research but not yet described in published literature, software must be made available to editors/reviewers
upon request. We strongly encourage code deposition in a community repository (e.g. GitHub). See the Nature Research guidelines for submitting code & software for further information.

Data
April 2018

Policy information about availability of data


All manuscripts must include a data availability statement. This statement should provide the following information, where applicable:
- Accession codes, unique identifiers, or web links for publicly available datasets
- A list of figures that have associated raw data
- A description of any restrictions on data availability
All data is available from the corresponding author upon reasonable request

1
nature research | reporting summary
Field-specific reporting
Please select the best fit for your research. If you are not sure, read the appropriate sections before making your selection.
Life sciences Behavioural & social sciences Ecological, evolutionary & environmental sciences
For a reference copy of the document with all sections, see nature.com/authors/policies/ReportingSummary-flat.pdf

Life sciences study design


All studies must disclose on these points even when the disclosure is negative.
Sample size 3 independent biological experiments were conducted, affording statistically significant means and standard deviations

Data exclusions There are no data exclusions.

Replication We have not attempted to replicate these cell viability studies

Randomization No in vivo studies were conducted. For in vitro studies, researcher A seeded cells, dosed cells, and determined cell viability; researcher B
irradiated the cells with test and control samples. Dosing was done in random rows in a fully seeded plate, and irradiation was done in
random rows of cells exposed or not exposed to nanorod samples.

Blinding No in vivo studies were conducted. For in vitro studies, experiments were double-blinded, as cells were blinded by default, and researcher A
was not informed of wells that were irradiated.

Reporting for specific materials, systems and methods

Materials & experimental systems Methods


n/a Involved in the study n/a Involved in the study
Unique biological materials ChIP-seq
Antibodies Flow cytometry
Eukaryotic cell lines MRI-based neuroimaging
Palaeontology
Animals and other organisms
Human research participants

April 2018

You might also like