Sei sulla pagina 1di 40

Chemical distinctions between Stradivaris maple and

modern tonewood
Hwan-Ching Taia,1,2, Guo-Chian Lia,1, Shing-Jong Huangb, Chang-Ruei Jhua, Jen-Hsuan Chunga, Bo Y. Wanga,
Chia-Shuo Hsua, Brigitte Brandmairc, Dai-Ting Chungd, Hao Ming Chena, and Jerry Chun Chung Chana
a
Department of Chemistry, National Taiwan University, Taipei 10617, Taiwan; bInstrumentation Center, National Taiwan University, Taipei 10617, Taiwan;
c
Private address, 86567 Hilgertshausen-Tandern, Germany; and dChimei Museum, Tainan 71755, Taiwan

Edited by Jerrold Meinwald, Cornell University, Ithaca, NY, and approved November 21, 2016 (received for review July 10, 2016)

Violins made by Antonio Stradivari are renowned for having been the wood supply or deforestation, or that Stradivari preferred denser
preferred instruments of many leading violinists for over two woods that grew slowly during the Maunder Minimum (5).
centuries. There have been long-standing questions about whether Elemental analyses of AS and DG maple specimens conducted
wood used by Stradivari possessed unique properties compared with by Nagyvary et al. (6) detected unusual minerals, which implicated
modern tonewood for violin making. Analyses of maple samples chemical treatments. They also reported severe degradation of the
removed from four Stradivari and a Guarneri instrument revealed lignocellulosehemicellulose matrix in violins through hydrolysis
highly distinct organic and inorganic compositions compared with and oxidation, which was attributed to chemical manipulation (7).
modern maples. By solid-state 13C NMR spectroscopy, we observed However, the conventional belief held by some violin restorers was
that about one-third of hemicellulose had decomposed after three that Cremonese maple plates appeared stiffer or more elastic than
centuries, accompanied by signs of lignin oxidation. No apparent modern counterparts, generally assessed by tapping and listening
changes in cellulose were detected by NMR and synchrotron X-ray (8). This apparent paradox motivated us to analyze the maples of
diffraction. By thermogravimetric analysis, historical maples exhibited five Cremonese instruments from three independent sources and

CHEMISTRY
reduced equilibrium moisture content. In differential scanning calorim- compare their organic fiber composition, cellulose crystallinity,
etry measurements, only maples from Stradivari violins, but not his moisture content, thermooxidative properties, and inorganic
cellos, exhibited unusual thermooxidation patterns distinct from nat- elemental composition.
ural wood. Elemental analyses by inductively coupled plasma mass The maple samples analyzed in this study are listed in Table 1
spectrometry suggested that Stradivaris maples were treated with (see SI Appendix, Table S1 for details). Four of our historical
complex mineral preservatives containing Al, Ca, Cu, Na, K, and Zn.
specimens were removed from Cremonese instruments during
This type of chemical seasoning was an unusual practice, unknown to
interior repairs of back plates. Three of them have been previously
later generations of violin makers. In their current state, maples in
analyzed by Nagyvary et al. (1717 AS violin; 1731 AS cello; 1741
Stradivari violins have very different chemical properties compared
DG violin) (6, 7), with the remaining materials transferred to us by
with their modern counterparts, likely due to the combined effects
courtesy of J. Nagyvary. The other back plate sample (1707 AS
of aging, chemical treatments, and vibrations. These findings may in-
spire further chemical experimentation with tonewood processing for
cello) was repaired at a different workshop (courtesy of Guy
instrument making in the 21st century.
Rabut). We also received the original neck of a 1725 AS violin
(courtesy of Chimei Museum). The neck was originally affixed to
the body with nails, which eventually rusted. It had to be modified
|
Cremona stringed instrument | wood treatment | wood aging |
Italian violin
Significance

T he invention of the modern violin has generally been attrib-


uted to Andrea Amati, who initiated the golden age of violin
making in Cremona, Italy (15501750). The most famous Cre-
There have been numerous attempts to elucidate the secrets of
Stradivari violins, to explain why functional replacements have not
monese violin maker was Antonio Stradivari (AS, or Antonius been reproduced over the past two centuries. Whether there are
systematic differences between Stradivari violins and later imita-
Stradivarius, 16441737), whose instruments are renowned for
tions has been heatedly debated. Our analysis of Stradivaris ma-
their tonal qualities. Stradivaris success in violin building was
ples from three independent sources showed reproducible
only matched by his neighbor, Giuseppe Guarneri del Ges differences in chemical compositions compared with modern ma-
(DG, or Joseph Guarnerius, 16981744), and leading violinists ples. Stradivaris use of mineral-treated maples belonged to a
today still mostly prefer to play AS and DG instruments. It re- forgotten tradition unknown to later violin makers. His maple also
mains a great mystery why functionally equivalent instruments appeared to be transformed by aging and vibration, resulting in a
have not been successfully constructed to replace these heavily unique composite material unavailable to modern makers. Modern
repaired Cremonese antiques, despite rapid technological advances chemical analyses may, therefore, improve our understanding of
over the last 200 y. For both historical and practical purposes, it is Stradivaris unique craft and inspire the development of novel
important to understand what types of materials were used to material approaches in instrument making.
construct these famous violins.
Violin acoustics is mainly determined by vibrations of the front Author contributions: H.-C.T., B.B., D.-T.C., H.M.C., and J.C.C.C. designed research; B.B. and
D.-T.C. provided historical consultation; G.-C.L., S.-J.H., C.-R.J., J.-H.C., B.Y.W., and C.-S.H.
and back plates, made of spruce and maple, respectively. It has performed research; G.-C.L., S.-J.H., C.-R.J., and J.-H.C. analyzed data; and H.-C.T. wrote
long been speculated that Stradivaris wood possessed unusual the paper.
properties not found in tonewood used by modern makers. Based The authors declare no conflict of interest.
on visual and microscopic inspections, experts suggested that This article is a PNAS Direct Submission.
wood varieties chosen by Stradivari have always been commer- Freely available online through the PNAS open access option.
cially available (13). Wood densities in Cremonese violins and 1
H.-C.T. and G.-C.L. contributed equally to this work.
modern tonewood also appeared to be similar, judging by com- 2
To whom correspondence should be addressed. Email: hctai@ntu.edu.tw.
puted tomography measurements (4). There is little evidence to This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
suggest that the Cremonese tradition was lost due to diminished 1073/pnas.1611253114/-/DCSupplemental.

www.pnas.org/cgi/doi/10.1073/pnas.1611253114 PNAS Early Edition | 1 of 6


Table 1. Maple samples analyzed in this study quantitative comparison was determined to be 1 ms, which was
Sample Description Origin also the conventional standard for studies on modern and ar-
chaeological woods (10, 11). Typical 13C{1H} CPMAS spectra
M1 Modern #1 European, flamed with contact time of 1 ms are shown in Fig. 1. The peak intensities
M2 Modern #2 European, flamed of these NMR spectra (SI Appendix, Figs. S12S14 and Table S2)
M3 Modern #3 European, flamed were subjected to principal component analysis (PCA), as shown
M4 Modern #4 European, plain in Fig. 2. We observed that historical maples were clearly sepa-
M5 Modern #5 European, plain rated from the modern controls along the first principal compo-
H1 1707 AS cello Back plate nent, which was primarily associated with the following peaks: 22,
H2 1717 AS violin Back plate, near the lower edge 56, 63, 72, 75, and 153 ppm. Differences along the first principal
H3 1725 AS violin neck Neck heel, original wood component (65% of the variance) were therefore strongly corre-
H4 1731 AS cello Back plate lated with hemicellulose changes, followed by lignin alterations.
H5 1741 DG violin Back plate, center region The second principal component further differentiated the his-
H6 Neck extension ca.1800 Heel extension on the 1725 neck torical neck samples from the historical back plate samples.
The rate of hemicellulose decomposition was estimated by the
hydrolysis of acetyl esters (22 ppm, acetyl CH3) and the chemical
by inserting new wood and reattached by gluing, which probably transformation of furanoses and pyranoses (C-2, C-3, and C-5 at
occurred around 1800 (2, 9). We cut into both the original and 75 ppm and C-6 at 63 ppm). After the subtraction of cellulose
extension parts to retrieve previously untouched wood from deep background signals, the rates of decomposition were found to be
locations (SI Appendix, Fig. S1). The extension piece (ca. 1800) similar at these three peaks (SI Appendix, Table S3). The average
provided a side-by-side aged control for Stradivaris neck material. signal reduction of hemicellulose was 36.9 3.9% (mean SE) in
The five modern controls were maple tonewoods recently pur- five Cremonese specimens compared with five modern controls (P <
chased in northern Italy. Our measurements revealed that Cre- 0.0001, Welchs t test with unequal variance). Modern and historical
monese maple clearly differed from modern tonewood with regard samples could be clearly differentiated by the extent of hemicellulose
to both organic and inorganic compositions, and the reasons for decomposition (Fig. 3A), providing experimental verification for the
such distinctions were further investigated and discussed. aged nature of our Cremonese samples. Assuming first-order ki-
netics, the half-life of hemicellulose decomposition was estimated
Results and Discussion around 409 56 y (mean SD) by regression analysis (SI Appendix,
Hemicellulose Decomposed in Historical Maples. The organic fiber Fig. S15). This was in reasonable agreement with previously pub-
compositions of historical and modern maples were analyzed lished measurements for dry-aged maple (358 y) (12), spruce (833
by 13C{1H} cross-polarization magic angle spinning (CPMAS) and 335 y) (12, 13), and fir (739 y, SI Appendix, Table S4) (14).
NMR spectroscopy. Historical maples exhibited reduced 13C
intensities at hemicellulose-related peaks (22, 62, 72, 75, and Lignin Became Oxidized in Historical Maples. In historical maples,
84 ppm) and lignin-related peaks (56 and 153 ppm) compared signal reductions at the methoxy (56 ppm) and aromatic carbons
with modern maples. These changes were evident across the (153 ppm) were indicative of lignin oxidation (SI Appendix, Fig.
spectra acquired at different CP contact times (0.2 ms to 10 ms, S14). There was only a weak correlation between lignin sig-
SI Appendix, Figs. S2S11), and the optimal contact time for nal changes at methoxy and aromatic carbons, suggesting that

Fig. 1. The 13C{1H} CPMAS spectra of maples de-


rived from modern control (M3) and three historical
specimens. The CP contact time was set to 1 ms. The
peaks were assigned to lignin, total cellulose (TC),
crystalline cellulose (CC), amorphous cellulose (AC),
hemicellulose (HC), and lignin.

2 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1611253114 Tai et al.


1717 AS violin and 1741 DG violin compared with modern
maples and 1731 AS cello, which was interpreted as evidence of
special chemical treatments applied to violins but not to cellos.
However, our 13C{1H} CPMAS data did not agree with their
previous findings. Our reanalysis of the same violin samples de-
tected only modest changes in hemicellulose and lignin, based on
extensive comparisons using different CP contact times (0.2 ms to
10 ms, SI Appendix, Fig. S11). We also applied the multiple-cross
polarization (multiCP) measurement method (SI Appendix, Fig.
S19 and Table S6), which has been recently proposed to improve
quantitation accuracy in 13C{1H} CPMAS NMR (20), to re-
confirm our observation. There were only minor spectral differ-
ences among the 1717 AS violin, 1725 AS violin neck, and 1731 AS
cello in multiCP measurements (SI Appendix, Fig. S19).
Two factors may have led to the discrepancy between our NMR
data and those of Nagyvary et al. (7). First, the maple shavings of
1717 and 1741 violins previously studied by Nagyvary et al. were
removed from interior areas that exhibited surface blackening,
possibly caused by surface heating. This discoloration was highly
Fig. 2. PCA score plot for the peak intensities of the CPMAS spectra mea- unusual for Cremonese instruments, and we were informed by
sured in maple samples. The first principal component (PC1) can clearly J. Nagyvary that the darker surface portions have been used up for
distinguish the modern and historical maples. Different colors represent analyses in the original reports (6, 7), whereas we only received
different sample origins (green for modern, red for historical back plates, residual samples from deeper regions (>0.3 mm from the surface)
and blue for historical necks).
showing normal coloration consistent with the yellowing of natu-

CHEMISTRY
rally aged wood (SI Appendix, Fig. S20). Second, Nagyvary et al.
multiple oxidation mechanisms may have been involved. There (7) used a fixed CP contact time of 5 ms for all of the NMR
were two plausible oxidation mechanisms in Cremonese speci- measurements, where it was implicitly assumed that all carbon
mens: oxidation by atmospheric oxygen and photooxidation. signals would exhibit the same decay rate during the contact time.
Stradivaris own letter indicated the exposure of instruments to By contrast, we carried out the CP experiments with different
sunlight (1), and further light exposures were inevitable over the contact times, and, indeed, we observed that there was a large
course of three centuries. Oxidation may cause lignin demethox- variation in signal decay rate for the carbon signals across different
ylation (15) and the formation of quinone groups (7, 16), resulting samples (SI Appendix, Fig. S11 and Table S7). Consequently, the
in peak broadening due to increased structural heterogeneity, as spectra acquired with the contact time of 5 ms were not warranted
well as decreased CP signals due to hydrogen loss nearby. In for quantitative analyses.
modern wood, the carbonyl peak at 173 ppm is mainly associated
Thermal Analyses of Historical Maples. Interestingly, a peculiar dif-
with the acetyl groups in hemicellulose. However, the carbonyl
signal at 173 ppm in historical maples did not show a corre- ference was found between Stradivaris violin and cello samples
sponding decrease to the acetyl-CH3 signal at 22 ppm (SI Ap- through differential scanning calorimetry (DSC) analyses (Fig. 4
pendix, Fig. S14). This was probably due to the formation of and SI Appendix, Table S8). When gradually heated to 600 C under
carbonyls and carboxylic acids from lignin oxidation (17), and air atmosphere, the thermograms of modern maples and two AS
plausibly from the oxidation of carbohydrates as well. cellos exhibited similar exothermic profiles with only two peaks. In
contrast, both 1717 AS violin and 1725 AS violin neck exhibited
Cellulose Remained Stable in Historical Maples. Cellulose microfibrils three exothermic peaks. The unique DSC profile associated with
in the wood contain individual rod-shaped crystalline domains sur- Stradivari violins has never been observed in natural wood, but only
rounded by amorphous regions. Because amorphous cellulose is less in fungus-degraded wood (21) and simple mixtures of holocellulose
stable than crystalline cellulose, the former may preferentially de- and lignin powders without mutual impregnation (22). It was noted
grade over time, leading to higher crystalline/amorphous cellulose that when holocellulose and lignin were mixed together by copre-
ratios (18). However, this was not observed in Cremonese maples, as cipitation to promote molecular adhesion, the thermogram
we estimated the relative degree of cellulose crystallinity by the in- only showed two peaks, just like natural wood (22). Hence, the
tensity ratio of the NMR peaks at 89 and 105 ppm (crystalline cel-
lulose/total cellulose) (11, 19), which appeared to be unaltered in
Cremonese samples (Fig. 3A). We also measured the lengths and
widths of cellulose crystalline domains by wide-angle X-ray scattering
using synchrotron radiation, based on the (004) and (200) reflections,
respectively (SI Appendix, Figs. S16S18). Crystallite lengths and
widths, as well as d spacing, have also remained largely invariant in
historical maples (Fig. 3B and SI Appendix, Table S5). We observed
neither the evidence for the degradation of amorphous cellulose, nor
the interconversion between crystalline and amorphous cellulose, nor
the changes in crystallite size and density. Our findings with respect to
cellulose stability, mild lignin oxidation, and significant hemicellulose
decomposition in historical maples were in reasonable agreement
with the reported effects of dry aging on wood (8).
Fig. 3. (A) Relative cellulose crystallinity plotted against relative hemi-
Quantitation of Cell Wall Composition by Different NMR Techniques. cellulose levels measured by 13C{1H} CPMAS. (B) Crystallite lengths and
Previous, Nagyvary et al. (7) have shown by 13C{1H} CPMAS widths estimated from X-ray diffraction patterns. Filled diamonds, modern
that hemicellulose and lignin signals were severely reduced in maples; open squares, historical specimens.

Tai et al. PNAS Early Edition | 3 of 6


to be detected by recent studies (28, 29). Modest increases (tens of
parts per million) in fungicidal elements including Cu, Zn, and B
may have been introduced by either surface application or liquid
infusion, but much greater increases in Ca, Na, K, and Al could
only be explained by the latter. In the 18th century, mineral in-
fusion could have been achieved by extended soaking and/or sap
displacement procedures using gravity flow (30). Both soaking and
sap displacement could help remove sap nutrients that support
microorganism growth and help introduce various preservatives
such as alum, sodium chloride, or vitriols (sulfates of copper
and zinc).
Because minerals initially introduced into the wood could
have been washed out during later treatment steps, the residual
elements currently detected in Stradivaris maples are insuffi-
cient for ascertaining the original mineral formulations or ap-
plication procedures. There is also no available method to detect
past heat treatments such as boiling, steaming, baking, or fuming.
It is therefore plausible that chemical or physical treatments may
account for the different DSC profiles between Stradivaris vio-
lins and cellos, but different vibration frequencies may also be
critical. Even a few hours of vibration has been shown to reduce
Fig. 4. DSC thermograms of modern maple (M3) and five historical
specimens.
the internal friction of wood, presumably by rearranging hydro-
gen bonds and polymer chains (31). Therefore, it is conceivable
that high-frequency vibrations could have gradually altered
molecular adhesion between holocellulose and lignin appeared to the ultrastructure of wood fibers, acting in concert with hemi-
have been reduced in Stradivari violins, even compared with his cellulose decomposition over time. The interactions among
cellos from the same era. This may help explain why Nagyvary et al. mineral preservatives, age-dependent chemical changes, me-
(7) observed obvious discrepancies between Stradivaris violins and chanical vibrations, and the composite nature of wood are extremely
cellos by 13C{1H} CPMAS NMR, as changes in wood fiber ultra- complex, and the potential outcomes are not well understood.
structure may affect the spin dynamics of the CP process. Although chemical changes of wood generally have measurable
By thermogravimetric analysis (TGA), we also observed a re- effects on its vibrational properties (8), the acoustic effects as-
duction of moisture content in Cremonese maples by about 25% sociated with the specific chemical alterations observed in this
(Fig. 5 and SI Appendix, Table S9). Four modern maples exhibited study remain to be investigated.
an average equilibrium moisture content (EMC) of 11.55 0.33%
(mean SE) at about 25 C and 58% relative humidity, in good Availability of Mineral-Treated Wood. Violin models designed by
agreement with reported values for maple (23). The four Stradi- Stradivari have been widely copied for over two centuries by
vari maples exhibited 8.67 0.53% EMC, and the difference was violin makers hoping to recreate similar playing qualities, but
highly significant (P < 0.01, Welchs t test with unequal variance). gaining access to wood materials resembling Stradivaris maple
Reduced hygroscopicity in historical maples may be attributed to may be difficult. First, it is unclear if there are accelerated aging
the decomposition of hemicellulose (24, 25), which is more hy- methods that can faithfully mimic the effects of natural aging.
groscopic than cellulose and lignin (26). There have been numerous anecdotes about artificial treatments
in modern instruments that led to wood structural damages and
Cremonese Maples Received Mineral Treatments. To understand irreparable failures after just 10 y to 20 y of playing. Secondly,
whether chemical treatments had possibly altered wood properties, wood infusion with mineral solutions was not a widespread
three historical samples (1731 cello, 1725 neck, and neck exten- practice and it has long been obsolete (32). Because violin plates
sion), with sufficient materials, were quantified for 25 elements are only a few millimeters thick, dimensional stability is an ex-
using inductively coupled plasma mass spectrometry (ICP-MS), tremely important tonewood quality. Immersion of dried tone-
with the results listed in SI Appendix, Table S10. These included 20 wood into liquids is likely to cause swelling damage and
metals and 5 metalloids/nonmetals (As, B, Ge, Sb, and P), but Cl, S, compromise dimensional stability, and is therefore generally
and Si could not be included due to instrumental limitations. In the
PCA analysis of the quantified elements (Fig. 6), the neck extension
and modern controls were closely clustered and represented wood
receiving little or no mineral treatment, with total metal content
around 2,000 ppm. By contrast, the total metal content was elevated
to 9,000 ppm in the 1725 neck and 1731 cello, which clearly stood
out as being unnatural in the PCA diagram. These two Stradivari
instruments showed similar increases in K, Na, Ca, Cu, and Zn. The
AS neck contained elevated levels of Al, and the AS cello was
higher in B.
Nagyvary et al. (6) have previously shown in a qualitative
manner, by electron probe microanalyzer, that maples used by
Stradivari and Guarneri had altered mineral composition. Here
we provide a quantitative assessment of elemental changes by
ICP-MS measurements. The changes detected in our experiments
could neither be simply be attributed to surface applications of Fig. 5. (A) TGA curves of modern and historical maples. (B) EMC values
potassium silicate (3) nor Pozzolana ash (27), which have been derived from TGA data (**P < 0.01 by Welchs t test with unequal variance; n =
previously proposed as wood sealers used by Stradivari but failed 4 for each group).

4 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1611253114 Tai et al.


Conclusion
In this study, we provided a quantitative assessment of wood
chemical changes in antique musical instruments, using a com-
bination of five analytical methods: NMR, synchrotron X-ray,
DSC, TGA, and ICP-MS. In Cremonese maples, we observed
normal chemical changes associated with dry aging, including
hemicellulose decomposition, lignin oxidation, and reduced
EMC. However, we also observed an unnatural elevation of
certain inorganic elements, which could be attributed to the use
of mineral-treated maples by Cremonese makers.
From a conservation perspective, it appears that maples in
Stradivari cellos are reasonably preserved. The crystallinity of
cellulose has remained intact, as well as the adhesion between
holocellulose and lignin. The mechanical strength of aged wood
depends on the stability of crystalline cellulose as well as the in-
tegrity of the surrounding lignincarbohydrate matrix (25, 40, 41).
Hence, it is a concern that continued decomposition of hemi-
cellulose may eventually cause structural breakdown.
In Stradivari violins, we already observe evidence of reduced
adhesion between holocellulose and lignin. It is plausible that high-
frequency vibration accelerated the breakdown of cell wall com-
Fig. 6. PCA score plot for elemental compositions measured in maples by ponents, but whether wood treatment may have also played a role is
ICP-MS. The 1731 cello (H4) was analyzed twice using two separate wood still unclear. It remains possible that some breakdown was initially
fragments. Different colors represent different sample origins (green for promoted by heat treatments or acidic/basic pH. However, it is also

CHEMISTRY
modern, red for historical back plates, and blue for historical necks).
possible that increased divalent and trivalent metal ions could cross-
link carbohydrate and lignin chains to compensate for reduced mo-
avoided. Before the introduction of pressurization methods in lecular adhesion. Additional antique samples and analytical methods
the 1800s, extensive mineral penetration by soaking or sap dis- are required to further elucidate the complex interactions between
placement was probably only achievable with water-saturated aging, wood treatments, and long-term vibrations, which appear to
have jointly transformed the chemical properties of maples in Stra-
wood, before full drying took placelikely applied by wood
divari violins. Changes in spruce soundboards also warrant further
suppliers rather than violin makers.
investigation. Our findings may inspire further chemical experimen-
To our knowledge, current supplies of commercial tonewood are
tation with wood for 21st century instrument making, applicable not
not mineral-infused, but just air-dried for 3 y to 10 y or even up to a
just to violins but also to wooden instruments across different cul-
century to achieve ideal dimensional stability (8). Mineral infusion
tures. Stradivaris maple may represent a singular case in the history
was not detected in our modern controls and the neck extension
of wooden musical instruments, but its implications and impact may
sample (Italy, ca. 1800), nor in the Jay viola (England, 1769) and
be far-reaching.
Gand & Bernardel violin (France, ca. 1845) previously analyzed by
Nagyvary et al. (6). Moreover, mineral treatment of wood has not Materials and Methods
been mentioned in published treatises on violin making that con- Wood Samples. Historical maple samples were taken from these instruments:
tain lutherie knowledge tracing back to the 18th and 19th centuries 1707 AS cello (H1), 1717 AS violin (H2), 1725 AS violin neck (H3), 1731 AS cello
(13, 33, 34). Building stringed instruments using mineral-treated (H4), and 1741 DG violin (H5), and neck extension ca. 1800 (H6). The modern
maples probably belonged to a long-forgotten tradition with few controls included five tonewood-grade maples of European origin (M1M5).
practitioners to begin with. To this date, we have not identified the See SI Appendix, Table S1 for details.
use of mineral-infused wood outside Cremona, which may partly
explain why some restorers have noticed that worm damages were NMR Spectroscopy. The 13C{1H} CPMAS NMR spectra were acquired at 13C and
1
relatively rare in Cremonese instruments. H frequencies of 150.9 MHz and 600.1 MHz, respectively, on a Bruker
Avance III NMR spectrometer (14.1 T) equipped with a commercial 4-mm
Nevertheless, wood preservation using mineral treatment had a
probe. Wood samples (10 mg) were finely cut using scalpels and spun at
long history in Italy and surrounding countries. Alum treatment of 12.5 kHz with recycle delay of 2.0 s. During the CP contact, the 1H nutation
wood was first recorded by ancient Romans for flame retardation frequency was linearly ramped from 35 kHz to 50 kHz, and that of 13C was
(35), and Paracelsus described, in the 16th century, the use of sal optimized for the HartmannHahn matching condition. Proton decoupling
gemma (alum) to enhance wood durability (36). Writings from of 75 kHz was applied during the t2 acquisition.
the 16th century about the positive effect of mineral treatments The number of scans was typically 8,192 for measurements with the
on instrument acoustics have been recently rediscovered (37). contact time of 1 ms and 4,096 for experiments with variable contact times
Chemical seasoning of wood using various minerals was still used (0.2 ms to 10 ms). Because of the limited resolution of the NMR spectra for
in the 20th century timber industry, but infrequently (32). With intact wood, we chose to analyze the peak intensities of the spectra without
performing spectral deconvolution. For the variable contact time experi-
our limited analytical data, it is difficult to elucidate the effects of
ments, the peak intensities were analyzed by the following equation (42):
mineral treatments in Stradivari instruments. For instance, we do
 
not know if the pH was sufficiently acidic or basic during the It = A1-exp-t=TCH exp -t T1 ,
treatment to promote hemicellulose hydrolysis. It is also possible
that Al and Ca ions added to the wood may act as chemical cross- where I(t) is the signal intensity at the contact time t, A is the intensity
linkersbeing chelated by hydroxyl, phenol, and carboxylate factor, TCH characterizes the time constant of the polarization transfer
between 1 H and 13 C, and T1 is the 1H relaxation time in the spin-locking
groups on hemicellulose, lignin, and cellulose (38, 39)to com-
radio frequency (RF) field. The curve fitting was performed in Origin2015
pensate for the natural degradation of hemicellulose over software (OriginLab).
time. Additional antique instruments from Cremona and For the multiCP measurements (20), the spinning frequencies were set at
other Italian cities need to be analyzed to elucidate if there 12 kHz. Typical /2 pulse lengths of 4 s and 3.57 s were applied for the 13C
were unique wood properties associated with Cremonese and 1H channels, respectively. Proton decoupling field strength of 75 kHz
makers or Stradivari himself. was used during the acquisition. A total of nine CP blocks were implemented

Tai et al. PNAS Early Edition | 5 of 6


with the contact time of 1 ms and the RF amplitude ramping of 90 to 100%. power and <5-W reflected power, with plasma/auxiliary/nebulizer flow rates
The recycle delay was 2 s, and the duration of the repolarization period tz at 15/0.8/0.95 L/min.
was 0.9 s. After the multiCP excitation, a Hahn echo of two rotor periods was
applied before the signal detection. Thermal Analysis. DSC and TGA experiments were conducted using a Mettler
TGA/DSC 1 instrument. Before analysis, finely cut wood samples were equilibrated
Wide-Angle X-Ray Scattering. The experiments were performed at the BL01C2 for 48 h in a sealed chamber with saturated sodium bromide solution, with room
beamline of the National Synchrotron Radiation Research Center, in which the temperature around 25 C and internal relative humidity around 58%. It only
ring was operated at 1.5 GeV energy with a typical current around 360 mA. A took around 5 min to transfer 5 mg of wood from the chamber to an alumi-
thin slice of wood sample was fixed with 3M Magic Tape and placed on the num crucible with a pierced lid and to have it stabilized inside the instrument at
sample stage. Two pairs of slits and one collimator were set up to provide a 34 C. Subsequently, the sample was heated at 10 C/min under air atmosphere
collimated beam with dimensions of 0.1 0.1 mm (H V) at the sample. The and recorded from 35 C to 600 C. EMC was calculated as (W35 W150)/W150,
wavelength of the incident X-rays was 1.033210 (12 keV), delivered from the where W35 and W150 were weights recorded at 35 C and 150 C, respectively.
5-T Superconducting Wavelength Shifter and an Si (111) triangular crystal
monochromator. The scattering signal was collected with a Mar345 imaging Color Photography. Small pieces of wood samples were placed on white filter
plate area detector, setting the exposure duration at 30 s. The diffraction papers inside a desktop photo studio with external lighting using white
angles were calibrated according to Bragg positions of CeO2 (SRM 674b) compact fluorescent lamps. A Takumar SMC 50 mm F1.4 lens (Ricoh) was
standards in desired geometry, and then GSAS II software (Argonne National mounted in reverse using appropriate adapters on a 24.3-megapixel Sony alpha
Laboratory) was used to obtain corresponding one-dimensional powder dif- ILCE-7 camera. The reversed connection resulted in a flange focal distance of
fraction profile with cake-type integration. The profile data were exported to 45.46 mm and ensured that all in-focus objects had the same magnification
Origin2015 software for baseline correction and smoothing. The size of the ratio. We performed color calibration on wood images by photographing
crystalline domain (L) was estimated using the Scherrer equation QPcard 203 and using QPcalibration software ver. 1.99b, with final processing in
Adobe Lightroom ver. 5.7.1.
L = K =B cos,

where K is the shape factor equal to 0.94 for wood cellulose (43), is the Statistical Tests. Welchs t test and linear regression were analyzed using
X-ray wavelength, B is the full width at half maximum in 2 units, and is Graphpad Prism software ver. 5.0. PCA was performed with Origin2015
the Bragg angle. software by choosing the correlation matrix (standardized) option.

ICP-MS. About 20 mg to 50 mg of finely cut wood sample was weighed and ACKNOWLEDGMENTS. We thank Joseph Nagyvary, Chimei Museum, and Guy
digested with 5 mL of 70% (wt/wt) nitric acid (trace metal grade) in a 50-mL glass Rabut for providing historical wood samples, and Sandro Chiao and Boa-Tsang
Lee for providing modern maples. We thank Wen-Feng Chang of the Instrument
centrifuge tube precleaned with 20% (wt/wt) nitric acid. The digestion mixture
Center at National Tsing-Hua University (Hsinchu, Taiwan) for ICP-MS measure-
was heated in a 100 C water bath for 8 h, followed by dilution with deionized ments, and the National Synchrotron Radiation Research Center (Hsinchu,
water to 50 mL, and analyzed by Agilent 7500ce ICP-MS. High-purity standards Taiwan) for technical services. This work was supported by Grant 102R890927
of individual elements (1,000 mg/L) were diluted in 1% nitric acid to from National Taiwan University, and Grants 103-2113-M-002-007-MY2 and 105-
prepare standards for calibration curves. ICP-MS was operated at 1.5-kW RF 2633-M-002-001 from Ministry of Science and Technology, Taiwan.

1. Hill WH, Hill AF, Hill AE (1902) Antonio Stradivari: His Life and Work (16441737) 21. Reh U, Kraepelin G, Lamprecht I (1986) Use of differential scanning calorimetry for
(Dover, New York). structural analysis of fungally degraded wood. Appl Environ Microbiol 52(5):11011106.
2. Pollens S (2010) Stradivari (Cambridge Univ Press, Cambridge, UK). 22. Tsujiyama S, Miyamori A (2000) Assignment of DSC thermograms of wood and its
3. Sacconi SF (1979) The Secrets of Stradivari, trans Dipper A, Rivaroli C (Libreria del components. Thermochim Acta 351(1-2):177181.
Convegno, Cremona, Italy). 23. Goulet M, Hernandez R (1991) Influence of moisture sorption on the strength of
4. Stoel BC, Borman TM, de Jongh R (2012) Wood densitometry in 17th and 18th century sugar maple wood in tangential tension. Wood Fiber Sci 23(2):197206.
Dutch, German, Austrian and French violins, compared to classical Cremonese and 24. Kohara J, Okamoto H (1955) Studies of Japanese old timbers. Sci Rep Saikyo Univ 7
modern violins. PLoS One 7(10):e46629. (1a):920.
5. Burcklem L, Grissino-Mayer HD (2003) Stradivari, violins, tree rings, and the Maunder 25. Yokoyama M, et al. (2009) Mechanical characteristics of aged Hinoki wood from
Minimum: a hypothesis. Dendrochronologia 21(1):4145. Japanese historical buildings. C R Phys 10(7):601611.
6. Nagyvary J, Guillemette RN, Spiegelman CH (2009) Mineral preservatives in the wood 26. Kollmann FFP, Cote WA, Jr (1968) Principles of Wood Science and Technology: I Solid
of Stradivari and Guarneri. PLoS One 4(1):e4245. Wood (Springer, Berlin).
7. Nagyvary J, DiVerdi JA, Owen NL, Tolley HD (2006) Wood used by Stradivari and 27. Barlow CY, Edwards PP, Millward GR, Raphael RA, Rubio DJ (1988) Wood treatment
Guarneri. Nature 444(7119):565. used in Cremonese instruments. Nature 332(6162):313.
8. Bucur V (2006) Acoustics of Wood (Springer, Berlin), 2nd Ed. 28. Echard JP, et al. (2010) The nature of the extraordinary finish of Stradivaris instru-
9. Hart G (1909) The Violin: Its Famous Makers and Their Imitators (Dulau, London). ments. Angew Chem Int Ed Engl 49(1):197201.
10. Bardet M, Foray MF, Maron S, Goncalves P, Tran QK (2004) Characterization of wood 29. Brandmair B, Greiner PS (2010) Stradivari Varnish (B. Brandmair, Munich).
30. Parnell EA (1844) Applied Chemistry: In Manufactures, Arts, and Domestic Economy
components of Portuguese medieval dugout canoes with high-resolution solid-state
(D. Appleton, New York).
NMR. Carbohydr Polym 57(4):419424.
31. Sobue N, Okayasu S (1992) Effects of continuous vibration on dynamic viscoelasticity
11. Gil AM, Pascoal Neto C (1999) Solid-state NMR studies of wood and other lignocel-
of wood. J Soc Mat Sci Jpn 41(461):164169.
lulosic materials. Annu Rep NMR Spectrosc 37:75117.
32. Kollmann FFP, Kuenzi EW, Stamm AJ (1975) Principles of Wood Science and
12. Pishik I, Fefilon V, Burkovskaya V (2011) Chemical composition and chemical prop-
Technology: II Wood Based Materials (Springer, Berlin).
erties of new and old wood. Lesnoi J 14(6):8993.
33. Regazzi R (1986) The Manuscript on Violin Making by G. A. Marchi - Bologna 1786,
13. Kranitz K (2014) Effect of natural aging on wood. Doctoral thesis (Eidgenssische
trans Sbarra N (A. Forni, Bologna, Italy).
Technische Hochschule Zrich, Zurich).
34. Roy K (2006) The Violin: Its History and Making (Karl Roy, Barrington, NH).
14. Kacik F, Smira P, Kacikova D, Reinprecht L, Nasswettrova A (2014) Chemical changes in
35. Cornell TJ (2013) The Fragments of the Roman Historians (Oxford Univ Press, Oxford).
fir wood from old buildings due to ageing. Cell Chem Technol 48(1-2):7988. 36. Waite AE (1894) The Hermetic and Alchemical Writings of Paracelsus (J. Elliott, London).
15. Paulsson M, Parkas J (2012) Review: Light-induced yellowing of lignocellulosic pulps 37. Gug R (1991) Salted soundboards and sweet sounds. Strad 102(1214):506511.
Mechanisms and preventive methods. BioResources 7(4):59956040. 38. Dheu-Andries ML, Perez S (1983) Geometrical features of calciumcarbohydrate in-
16. Ganne-Chedeville C, Jaaskelainen A-S, Froidevaux J, Hughes M, Navi P (2012) Natural teractions. Carbohydr Res 124(2):324332.
and artificial ageing of spruce wood as observed by FTIR-ATR and UVRR spectroscopy. 39. Nolte H, John S, Smidsrod O, Stokke BT (1992) Gelation of xanthan with trivalent
Holzforschung 66(2):163170. metal ions. Carbohydr Polym 18(4):243251.
17. Evstigneeva EI, et al. (2015) Chemical structure and physicochemical properties of 40. Lionetto F, Del Sole R, Cannoletta D, Maffezzoli A (2012) Monitoring wood degra-
oxidized hydrolysis lignin. Russ J Appl Chem 88(8):12951303. dation during weathering by cellulose crystallinity. Materials (Basel) 5(10):19101922.
18. Yonenobu H, Tsuchikawa S (2003) Near-infrared spectroscopic comparison of antique 41. Obataya E (2012) Effects of aging on the vibrational properties of wood. J Cult Herit
and modern wood. Appl Spectrosc 57(11):14511453. 13(3 Suppl):S21S25.
19. Wikberg H (2004) Advanced solid state NMR spectroscopic techniques in the study of 42. Love GD, Snape CE, Jarvis MC (1992) Determination of the aromatic lignin content in
thermally modified wood. Doctoral thesis (Univ Helsinki, Helsinki). oak wood by quantitative solid-state 13C-NMR. Biopolymers 32(9):11871192.
20. Johnson RL, Schmidt-Rohr K (2014) Quantitative solid-state 13C NMR with signal en- 43. Fernandes AN, et al. (2011) Nanostructure of cellulose microfibrils in spruce wood.
hancement by multiple cross polarization. J Magn Reson 239:4449. Proc Natl Acad Sci USA 108(47):E1195E1203.

6 of 6 | www.pnas.org/cgi/doi/10.1073/pnas.1611253114 Tai et al.


Supporting Information Appendix

for

Chemical distinctions between Stradivaris maple and modern tonewood

Hwan-Ching Tai1, Guo-Chian Li1, Shing-Jong Huang2, Chang-Ruei Jhu1, Jen-Hsuan Chung1,

Bo Y. Wang1, Chia-Shuo Hsu1, Brigitte Brandmair3, Dai-Ting Chung4, Hao Ming Chen1,

Jerry Chun Chung Chan1

1
Department of Chemistry, National Taiwan University, Taipei 10617, Taiwan;
2
Instrumentation Center, National Taiwan University, Taipei 10617, Taiwan;
3
Private address, 86567 Hilgertshausen-Tandern, Germany;
4
Chimei Museum, Tainan 71755, Taiwan


These authors contributed equally to this work

1
(I) Supplementary Tables

Table S1. Origins of maple samples analyzed in this study

Sampling depth
Sample Description Origin Provider
from surface

M1 modern #1 European [1] >2 mm Sandro Chiao

M2 modern #2 European [1] >2 mm Sandro Chiao

M3 modern #3 European [1] >2 mm Sandro Chiao

M4 modern #4 European [2] >2 mm Boa-Tsang Lee

M5 modern #5 European [2] >2 mm Boa-Tsang Lee

H1 1707 AS cello back plate [3] not available Guy Rabut

H2 1717 AS violin back plate, near the lower edge [4] >0.3 mm Joseph Nagyvary

H3 1725 AS neck neck heel, original wood [5] > 3 mm Chimei Museum

H4 1731 AS cello back plate [4] >0.3 mm Joseph Nagyvary

H5 1741 DG violin back plate, center [4] >0.3 mm Joseph Nagyvary

H6 neck extension c.1800 heel extension on the 1725 neck [6] > 3 mm Chimei Museum

Notes:
[1] Flamed tonewood from Sandro Chiaos workshop in Taipei, Taiwan. Purchased in Italy, circa 2005,
probably of Balkan origin. Maple species traditionally used by Italian violin makers included Acer
pseudoplatanus, Acer platanoides, and Acer campestre, the woods of which were practically
indistinguishable.
[2] Plain-figured tonewood from Boa-Tsang Lees workshop in Taipei, Taiwan. Purchased in Italy, circa
1995, probably of Italian origin.
[3] The sample was taken in the course of restoration in the 1980's and generously provided by Guy Rabut
(New York, NY).
[4] The sample was taken in the course of restoration and generously provided by Ren A. Morel (New
York, NY). Portions of the sample had been used for previous studies (1, 2), and the remaining
material was generously provided by Joseph Nagyvary.
[5] The original neck removed from the 1725 Stradivari Brancaccio violin, which was owned by the
Brancaccio family in Napoli during the 19th century. See Fig. S1 for sampling location.
[6] Wood inserted by an unidentified Italian restorer, circa 1800. See Fig. S1 for sampling location.

2
Table S2. 13C{1H} CPMAS NMR peak intensities of modern and historical maples

13
C chemical shift (ppm)
Sample
22 56 63 65 72 75 84 89 105 137 148 153 173

modern M1 0.307 0.691 0.850 0.749 2.158 2.379 0.494 0.315 1.000 0.085 0.078 0.201 0.086

modern M2 0.326 0.639 0.780 0.700 2.030 2.278 0.479 0.333 1.000 0.098 0.089 0.197 0.086

modern M3 0.331 0.628 0.794 0.724 2.064 2.265 0.446 0.318 1.000 0.087 0.083 0.185 0.084

modern M4 0.261 0.622 0.759 0.717 2.005 2.177 0.446 0.350 1.000 0.082 0.084 0.167 0.076

modern M5 0.313 0.575 0.764 0.695 1.979 2.249 0.464 0.345 1.000 0.085 0.095 0.183 0.085
neck extension
0.209 0.482 0.635 0.649 1.809 1.828 0.429 0.371 1.000 0.081 0.096 0.119 0.118
c. 1800
AS neck 1725 0.154 0.455 0.547 0.607 1.740 1.825 0.430 0.391 1.000 0.081 0.094 0.154 0.083

AS violin 1717 0.190 0.485 0.608 0.682 1.886 1.857 0.388 0.355 1.000 0.070 0.090 0.129 0.062

DG violin 1741 0.166 0.427 0.622 0.696 1.868 1.801 0.341 0.315 1.000 0.050 0.091 0.139 0.060

AS cello 1707 0.236 0.510 0.653 0.714 1.914 1.886 0.417 0.377 1.000 0.084 0.085 0.156 0.066

AS cello 1731 0.241 0.478 0.643 0.688 1.872 1.878 0.394 0.316 1.000 0.056 0.085 0.140 0.080

Note: CP contact time = 1 ms; the peak intensities were normalized with reference to the peak at 105 ppm
(C-1, total cellulose)

3
Table S3. Peak intensities of hemicellulose extracted from 13C{1H} CPMAS spectra

Mean
22 ppm 63 ppm 75 ppm
Sample (hemicellulose
(acetyl CH3) (C-6) (C-2, 3, 5)
level)
modern M1 0.997 1.125 1.100 1.074

modern M2 1.061 0.980 1.008 1.016

modern M3 1.075 1.010 0.996 1.027

modern M4 0.848 0.937 0.915 0.900

modern M5 1.018 0.947 0.981 0.982

neck extension c. 1800 0.679 0.680 0.593 0.651

AS neck 1725 0.501 0.498 0.591 0.530

AS violin 1717 0.618 0.624 0.620 0.621

DG violin 1741 0.540 0.654 0.568 0.587

AS cello 1707 0.769 0.718 0.646 0.711

AS cello 1731 0.782 0.697 0.639 0.706

Note: From Table S2, the cellulose background of the peaks at 63 and 75 ppm were estimated to
be 0.306 and 1.184, respectively. These values were derived from the spin-locking cellulose
spectra of Kraft pulp reported by Liitia (3). Note that the peak at 22 ppm does not have cellulose
background. After subtracting the cellulose background, the mean of each NMR peak of the five
modern maples was used to scale the intensities of the corresponding NMR peaks of the historical
maples.

4
Table S4. Linear regression analysis of hemicellulose decomposition rate

Age range Slope* Half-life Std. dev. of


Sample series R2 Data source
(year) (103 year1) (year) half-life

Tonewood
10308 1.693 0.8564 409 56 Table S3
maples

Maple 0100 1.934 0.9976 358 18 Pishik et al.#

Spruce 0700 0.8318 0.9605 833 76 Pishik et al.#

Spruce 0210 2.068 0.5947 335 160 Kranitz(4)

Fir 0389 0.9379 0.9773 739 56 Kacik et al.(5)

* Slope of log [hemicellulose/cellulose] vs. sample age


#
The original study by Pishik et al. was published in Russian (6), and the relevant data has been
summarized by Bucur (7).

5
Table S5. Wide-angle X-ray scattering data of modern and historical maples

(200) (004)
Sample
domain domain
2 FWHM* d (nm) 2 FWHM* d (nm)
size (nm) size (nm)
modern M1 14.85 1.763 3.18 0.400 23.10 0.1840 30.87 0.258

modern M2 14.83 1.871 3.00 0.400 23.08 0.1687 33.67 0.258

modern M3 14.82 1.871 3.00 0.401 23.08 0.1687 33.67 0.258

modern M4 14.91 1.858 3.02 0.398 23.14 0.2116 26.84 0.258

modern M5 14.91 1.822 3.08 0.398 23.10 0.2024 28.06 0.258

AS cello 1707 14.94 1.822 3.08 0.397 23.17 0.1748 32.50 0.257

AS cello 1731 14.76 1.886 2.98 0.402 22.97 0.2024 28.06 0.259

AS violin 1717 14.82 2.116 2.65 0.400 23.00 0.2116 26.84 0.259

DG violin 1741 14.84 1.674 3.35 0.400 22.98 0.1840 30.86 0.259

AS neck 1725 15.12 1.757 3.19 0.393 23.12 0.1932 29.40 0.258
neck extension
14.99 1.803 3.11 0.396 23.13 0.2024 28.06 0.258
c.1800

* full-width-at-half-maxima in 2 units (degrees)

6
Table S6. Comparison of the peak intensities between the 13C{1H} CPMAS spectrum
with contact time of 1 ms and the 13C{1H} multiCP spectrum

Neck extension
Peak Modern M3 AS neck 1725 AS violin 1717 AS cello 1731
c. 1800

(ppm) 1 ms multiCP 1 ms multiCP 1 ms multiCP 1 ms multiCP 1 ms multiCP

22 0.331 0.253 0.209 0.172 0.154 0.140 0.190 0.149 0.241 0.184

56 0.628 0.512 0.482 0.411 0.455 0.459 0.485 0.444 0.478 0.473

63 0.794 0.633 0.635 0.567 0.547 0.534 0.608 0.585 0.643 0.601

65 0.724 0.702 0.649 0.639 0.607 0.632 0.682 0.743 0.688 0.686

72 2.064 1.931 1.809 1.810 1.740 1.740 1.886 1.841 1.872 1.777

75 2.265 1.933 1.828 1.871 1.825 1.825 1.857 1.757 1.878 1.691

84 0.446 0.521 0.429 0.448 0.430 0.467 0.388 0.489 0.394 0.436

89 0.318 0.384 0.371 0.357 0.391 0.381 0.355 0.367 0.316 0.362

105 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000 1.000

137 0.087 0.208 0.081 0.119 0.081 0.153 0.070 0.148 0.056 0.133

148 0.083 0.171 0.096 0.156 0.094 0.129 0.090 0.183 0.085 0.169

153 0.185 0.373 0.119 0.290 0.154 0.327 0.129 0.269 0.140 0.272

173 0.084 0.138 0.118 0.121 0.083 0.211 0.062 0.120 0.080 0.138

7
Table S7. T1 data of modern and historical maples

13
C chemical shift (ppm)
Sample
22 56 75 84 89 105 153 173

T1 (ms) 9.54 9.34 6.87 6.61 7.40 7.33 16.93 17.12


modern M1
Std. dev. 0.71 0.64 0.20 0.22 0.36 0.26 2.91 6.12

T1 (ms) 6.21 6.57 4.02 3.48 3.89 4.47 6.58 7.29


modern M4
Std. dev. 0.42 0.39 0.22 0.29 0.27 0.16 0.51 1.02

T1 (ms) 6.91 6.12 4.19 4.10 4.79 4.86 7.79 6.33


modern M5
Std. dev. 0.65 0.35 0.33 0.42 0.45 0.20 0.95 1.82

T1 (ms) 6.47 6.42 4.27 3.84 4.37 4.86 6.74 8.30


AS cello 1707
Std. dev. 0.53 0.68 0.20 0.40 0.43 0.13 1.39 1.99

T1 (ms) 8.84 8.25 6.53 6.45 7.26 6.70 10.48 6.36


AS cello 1731
Std. dev. 1.11 0.32 0.22 0.45 0.59 0.17 1.31 1.04

T1 (ms) 10.72 8.71 7.28 7.52 8.30 7.40 10.32 18.49


AS violin 1717
Std. dev. 1.68 0.61 0.40 0.76 1.27 0.38 1.59 10.64

T1 (ms) 5.93 5.45 3.83 3.19 3.35 4.02 7.59 8.22


DG violin 1741
Std. dev. 1.13 0.34 0.12 0.40 0.30 0.08 1.94 1.91

T1 (ms) 9.08 8.30 6.48 7.06 7.44 7.11 9.79 4.80


AS neck 1725
Std. dev. 1.16 0.92 0.39 0.56 0.56 0.42 2.04 1.79

T1 (ms) 7.26 7.94 6.05 6.00 7.20 6.41 6.81 4.51


neck extension
c.1800
Std. dev. 1.14 0.65 0.29 0.64 0.59 0.12 1.95 1.89

8
Table S8. DSC peak profiles of modern and historical maples

Early peak Intermediary Late peak


Sample Sample number
(oC) peak (oC) (oC)
modern M1 355 ND 476
modern M2 359 ND 486
modern M3 361 ND 491
modern M4 346 ND 462
modern M5 364 ND 493
#1 344 ND 486
neck extension
c. 1800 #2 344 ND 492
AS cello 1707 358 ND 481
#1 347 ND 479
AS cello 1731
#2 350 ND 490
left side #1 317 419 464

AS violin neck left side #2 317 424 471


1725 right side #1 314 412 463
right side #2 314 417 469
#1 314 409 465
AS violin 1717
#2 315 408 463

ND = not detected

9
Table S9. Moisture content of modern and historical maples measured by TGA

Equilibration condition*
Weight at Weight at EMC
Sample
temperature relative 35 oC (mg) 150 oC (mg) (%)
(oC) humidity (%)

modern M1 25 58 6.0986 5.5147 10.6

modern M2 25 58 6.5295 5.8435 11.7

modern M3 25 58 5.6514 5.0547 11.8

modern M5 25 58 5.3106 4.7391 12.1

AS cello 1707 25 58 6.3026 5.8451 7.8

AS cello 1731 25 58 5.8127 5.3837 8.0

AS violin 1717 25 58 5.0166 4.6135 8.7

AS neck 1725 25 58 5.2546 4.7702 10.2

neck extension
25 58 5.1567 4.7421 8.7
c. 1800

* Equilibrated for 48 hours in a sealed chamber with saturated NaBr solution. The room temperature was
controlled between 2427 oC, which corresponded to 5758% relative humidity inside the chamber.

10
Table S10. ICP-MS measurements of modern and historical maples

neck 1731 AS 1731 AS


modern modern modern modern modern 1725 AS Detection
Element Unit extension cello cello
M1 M2 M3 M4 M5 neck limit
c. 1800 sample 1* sample 2*

Al ppm 10.6 289.6 114.5 2.6 52.0 ND 547.2 29.8 32.0 0.127
As ppm ND ND ND ND ND ND ND ND ND 0.002
B ppm 4.3 3.5 3.3 8.5 8.2 6.3 6.0 31.4 32.4 0.157
Ba ppm 0.6 0.9 0.6 1.5 0.8 1.3 1.2 2.1 2.9 0.005
Ca ppm 420.0 395.9 357.0 460.6 455.5 404.2 1448.0 1621.0 1867.4 0.332
Cr ppm ND ND ND ND ND ND ND ND ND 0.013
Cu ppm 0.9 1.6 1.3 1.4 1.0 4.6 15.8 30.0 33.6 0.007
Fe ppm 3.8 ND ND 2.1 ND ND 553.7 #
ND ND 0.356
Ge ppm ND ND ND ND ND ND ND ND ND 0.002
Hg ppm ND ND ND ND ND ND ND ND ND 0.098
K ppm 838.9 623.9 593.8 1789.4 1045.9 1035.4 4181.0 3709.0 3984.0 1.514
Li ppm ND ND ND ND ND ND ND ND ND 0.017
Mg ppm 302.8 405.1 403.8 450.5 262.6 449.0 271.8 183.2 214.4 0.027
Mn ppm 3.7 2.2 1.8 5.4 3.4 1.5 5.4 3.2 4.9 0.001
Na ppm 24.5 45.5 45.3 35.1 27.6 216.2 2727.0 3292.0 3485.0 0.529
Ni ppm ND 0.2 ND 0.3 0.5 0.2 1.1 10.2 12.8 0.004
P ppm 69.4 110.2 23.3 124.4 131.9 126.2 30.3 123.2 141.6 0.679
Pb ppm 0.3 0.3 0.2 0.2 0.3 2.9 1.6 7.7 10.5 0.003
Sb ppm ND ND ND ND ND ND ND ND ND 0.004
Se ppm ND ND ND ND ND ND ND ND ND 0.098
Sn ppm 0.3 0.3 0.2 0.4 0.3 0.4 1.1 1.1 1.1 0.003
Sr ppm 3.1 0.9 0.8 2.8 0.9 8.3 8.8 4.2 4.8 0.003
Ti ppm 2.5 19.2 3.3 3.6 6.6 4.8 1.5 27.7 35.3 0.001
Zn ppm 4.3 8.2 6.1 4.7 4.9 55.9 36.4 44.7 42.0 0.030
Zr ppm ND ND ND ND ND ND ND ND ND 0.001

Total
ppm 1616 1794 1529 2761 1862 2185 9248 8966 9731
metal

Notes:
* Two different wood fragments from the 1731 cello were analyzed independently
#
Possible Fe contamination by rusted nails inside
ND = not detected

11
(II) Supplementary Figure Legends

Figure S1. Photographs of the 1725 AS violin neck. (a) The original neck of the 1725 violin was
initially nailed to the body by Stradivari (black arrowhead indicates nail holes). (b) The open
arrowhead indicates where the inserted heel extension wood (c. 1800) was sampled for analysis.
Stradivaris original wood was sampled from both the left side and the right side (black arrowheads)
of the extension material, using a combination of mechanical drill and hand tools to retrieve wood
more than 3 mm beneath the surface.

Figure S2. 13C{1H} CPMAS NMR spectra of modern maple M1 with different contact times. The
experimental contact time was indicated for each trace.

Figure S3. 13C{1H} CPMAS NMR spectra of modern maple M4 with different contact times.

Figure S4. 13C{1H} CPMAS NMR spectra of modern maple M5 with different contact times.

Figure S5. 13C{1H} CPMAS NMR spectra of 1707 AS cello maple with different contact times.

Figure S6. 13C{1H} CPMAS NMR spectra of 1717 AS violin maple with different contact times.

Figure S7. 13C{1H} CPMAS NMR spectra of 1725 AS neck maple with different contact times.

Figure S8. 13C{1H} CPMAS NMR spectra of 1731 AS cello maple with different contact times.

Figure S9. 13C{1H} CPMAS NMR spectra of 1741 DG violin maple with different contact times.

Figure S10 13C{1H} CPMAS NMR spectra of neck extension (c. 1800) with different contact times.

Figure S11. Variation of 13C{1H} CPMAS peak intensities at different contact times. For each maple

12
sample, the peak intensities at different chemical shifts and different contact times from the same
sample were normalized with reference to the signal at 105 ppm with contact time equal to 1 ms.

Figure S12. 13C{1H} CPMAS NMR spectra of 5 modern and 6 historical maples. Acquired with 1
ms CP contact time and normalized to the total cellulose peak at 105 ppm.

Figure S13. 13C{1H} CPMAS NMR spectra of 5 modern and 6 historical maples, enlarged insets.
The spectra from Figure S12 are shown as enlarged insets for detailed comparisons. The peaks were
assigned to total cellulose (TC), crystalline cellulose (CC), amorphous cellulose (AC), hemicellulose
(HC), and lignin.

Fig. S14. Comparison of 13C{1H} CPMAS peak intensities of the spectra of modern and historical
maples. For brevity, only the mean of 5 modern maples are shown (Table S2), where the error bar
represents 95% confidence interval of the mean. Values of the historical maples were normalized
with reference to the mean of the modern maples.

Figure S15. Kinetics of hemicellulose decomposition in historical maples. Linear regression analysis
based on the mean hemicellulose levels in Table S3.

Figure S16. Wide-angle X-ray scattering patterns of selected maple specimens.

Figure S17. Diffraction peaks of (004) obtained for modern and historical maple samples.

Figure S18. Diffraction peaks of (200) obtained for modern and historical maple samples.

Figure S19. 13C{1H} multiCP spectra of maples from modern control (M3) and three historical
specimens.

Figure S20. Color comparison between modern and historical maple specimens.

13
(III) References

1. Nagyvary J, DiVerdi JA, Owen NL, & Tolley HD (2006) Wood used by Stradivari and
Guarneri. Nature 444:565.
2. Nagyvary J, Guillemette RN, & Spiegelman CH (2009) Mineral preservatives in the wood of
Stradivari and Guarneri. PloS one 4:e4245.
3. Liitia T (2002) Application of Modern NMR Spectroscopic Techniques to Structural Studies
of Wood and Pulp Components. Doctoral thesis (University of Helsinki, Helsinky, Finland).
4. Kranitz K (2014) Effect of natural aging on wood. Doctoral thesis (Eidgenossische
Technische Hochschule Zurich, Zurich, Switzerland).
5. Kacik F, Smira P, Kacikova D, Reinprecht L, & Nasswettrova A (2014) Chemical changes in
fir wood from old buildings due to ageing. Cellulose Chem. Technol. 48:79-88.
6. Pishik I, Fefilon V, & Burkovskaya V (2011) Chemical Composition and Chemical Properties
of New and Old Wood. Lesnoi J. 14:89-93.
7. Bucur V (2006) Acoustics of Wood (Springer, Berlin, Germany) 2nd Ed.

14
Figure S1

a b
nail holes

heel
extension
original
original
wood
wood
Figure S2
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S3
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S4
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S5
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S6
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S7
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S8
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S9
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S10
10.0 ms
7.5 ms
5.0 ms
3.0 ms
2.0 ms
1.5 ms
1.0 ms
0.75 ms
0.5 ms
0.2 ms
200 180 160 140 120 100 80 60 40 20 0
13
C chemical shift (ppm)
Figure S11

0.8

signal intensity at 56 ppm


signal intensity at 22 ppm

0.3
0.6

0.2
0.4

0.1
0.2

0.0 0.0
0.25 0.5 1 2 4 8 0.25 0.5 1 2 4 8
contact time (ms) contact time (ms)
3
signal intensity at 75 ppm

signal intensity at 84 ppm


0.6

2
0.4

1
0.2

0 0.0
0.25 0.5 1 2 4 8 0.25 0.5 1 2 4 8
contact time (ms) contact time (ms)
0.5 1.5
signal intensity at 105 ppm
signal intensity at 89 ppm

0.4

1.0
0.3

0.2
0.5

0.1

0.0 0.0
0.25 0.5 1 2 4 8 0.25 0.5 1 2 4 8
contact time (ms) contact time (ms)
0.30 0.16
signal intensity at 153 ppm

signal intensity at 173 ppm

0.25
0.12
0.20

0.15 0.08

0.10
0.04
0.05

0.00 0.00
0.25 0.5 1 2 4 8 0.25 0.5 1 2 4 8
contact time (ms) contact time (ms)

AS cello 1707 AS violin 1717 AS neck 1725


AS cello 1731 DG violin 1741 neck extension
modern M1 modern M4 modern M5
Figure S12

modern M1
modern M2
modern M3
modern M4
modern M5
neck extension
AS neck 1725
AS cello 1707
AS cello 1731
AS violin 1717
DG violin 1741
intensity

200 180 160 140 120 100 80 60 40 20 0


13
C chemical shift (ppm)
Figure S13
AC+HC
HC
CC lignin

30 28 26 24 22 20 18 16 14 12 10 70 68 66 64 62 60 58 56 54 52 50
13 13
C chemical shift (ppm) C chemical shift (ppm)

TC+HC AC+HC

CC

80 78 76 74 72 70 94 92 90 88 86 84 82 80
13 13
C chemical shift (ppm) C chemical shift (ppm)

lignin modern M1
modern M2
modern M3
modern M4
lignin
modern M5
neck extension
AS neck 1725
AS cello 1707
AS cello 1731

160 158 156 154 152 150 148 146 144 142 140 AS violin 1717
13
C chemical shift (ppm) DG violin 1741
Figure S14

modern (mean) DG violin 1741 AS violin 1717 AS cello 1731


AS cello 1707 AS neck 1725 neck extension
1.6

1.4
normalized intensity

1.2

1.0

0.8

0.6

0.4

0.2

0.0
173 153 148 137 105 89 84 75 72 65 63 56 22
13
C chemical shift (ppm)
Figure S15

0.2
log (hemicellulose level)

0.1

-0.1

-0.2

-0.3 H4 H1

-0.4 Y 2= -1.693E-03 X+1.56E-02


R = 0.8564 H2
-0.5 H6

-0.6 H5
H3
-0.7
0 50 100 150 200 250 300 350

Year
Figure S16

modern maple M3

AS neck 1725
Figure S17

maple M1
maple M2
maple M3
(004) maple M4
maple M5
neck extension
AS neck 1725
intensity

AS cello 1707
AS cello 1731
AS violin 1717
DG violin 1741

22 23 24

2 (degree)
Figure S18

maple M1
maple M2
maple M3 (200)
maple M4
maple M5
neck extension
AS neck 1725
intensity

AS cello 1707
AS cello 1731
AS violin 1717
DG violin 1741

8 10 12 14 16 18
2 (degree)
Figure S19

modern maple
AS neck 1725
AS violin 1717
AS cello 1731
intensity

180 170 160 150 140 130 120 30 25 20 15

200 180 160 140 120 100 80 60 40 20 0


13
C chemical shift (ppm)
Figure S20
modern maple

modern M1 modern M2 modern M3

5 mm

modern M4 modern M5

historical maple

AS cello 1707 AS cello 1731 AS violin 1717

5 mm
AS neck, 1725 DG violin 1741 neck extension
c. 1800

Potrebbero piacerti anche