100% found this document useful (1 vote)
1K views448 pages

Handbook of Colorants Chemistry 2

Uploaded by

Amer Kasideh
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
1K views448 pages

Handbook of Colorants Chemistry 2

Uploaded by

Amer Kasideh
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd

Ingo Klöckl

Handbook of Colorants Chemistry


Also of Interest
Handbook of Colorants Chemistry
Volume 1: Dyes and Pigments Fundamentals
Klöckl, 2023
ISBN 978-3-11-077699-7, e-ISBN 978-3-11-077711-6

Chemistry for Archaeology


Heritage Sciences
Reiche, Alfeld, Radtke, Hodgkinson, 2020
ISBN 978-3-11-044214-4, e-ISBN 978-3-11-044216-8

Chemical Analysis in Cultural Heritage


Sabbatini, van der Werf (Eds.), 2020
ISBN 978-3-11-045641-7, e-ISBN 978-3-11-045753-7

Encyclopedia of Color, Dyes, Pigments


Pfaff, 2022
Volume 1. Antraquinonoid Pigments – Color Fundamentals
ISBN 978-3-11-058588-9, e-ISBN 978-3-11-058807-1
Volume 2. Color Measurement – Metal Effect Pigments
ISBN 978-3-11-058684-8, e-ISBN 978-3-11-058710-4
Volume 3. Mixed metal Oxide Pigments – Zinc Sulfide Pigments
ISBN 978-3-11-058686-2, e-ISBN 978-3-11-058712-8
Ingo Klöckl

Handbook of
Colorants
Chemistry

Volume 2: in Painting, Art and Inks
Author
Dr. rer. nat. Ingo Klöckl
St. Leoner Strasse 16
68809 Neulußheim
Germany
[Link]@[Link]

ISBN 978-3-11-077700-0
e-ISBN (PDF) 978-3-11-077712-3
e-ISBN (EPUB) 978-3-11-077729-1

Library of Congress Control Number: 2022949958

Bibliographic information published by the Deutsche Nationalbibliothek


The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie;
detailed bibliographic data are available on the Internet at [Link]

© 2023 Walter de Gruyter GmbH, Berlin/Boston


Cover image: nikamata / iStock / Getty Images Plus
Typesetting: VTeX UAB, Lithuania
Printing and binding: CPI books GmbH, Leck

[Link]

For all those who, through curiosity, have discovered or may discover their interest in
the fascinating field of chemistry
Foreword
Nothing shows us the beauty of our world as vividly as its colors. For our distant bio-
logical ancestors, there were predominantly smells. However, at some point in our long
evolution into modern humans, we dared to swap the dull magic realm of scents for the
bright precision of our eyes. And yet colors are also a magical realm that holds many a
secret. Unlike shape, density, or surface texture, color is not an inherent property of an
object but only our perception of how the object reflects or absorbs visible light. More-
over, our eyes show us only a tiny fraction of the immense spectrum of electromagnetic
radiation that fills our universe. The wavelengths, and thus the frequencies, of this spec-
trum span 16 orders of magnitude—from the 10 to 20 km long radio waves of some mili-
tary transmitters to the gamma rays of imploding galaxies, which are only a thousandth
of a nanometer short. Life on our planet mainly registers wavelengths between 300 and
1000 nm. This range includes the ultraviolet, with wavelengths below 400 nm, which un-
like many insects, we cannot see; the range from blue to green to red, that is, from 400
to about 750 nm, which means light to us; and finally, infrared rays, with wavelengths
above 800 nm, which some animals perceive as light but we perceive only as heat.
Ultraviolet was probably the first color that life on our planet saw. This spectral part
of sunlight meant danger, as it destroyed many biological building blocks. Cells devel-
oped a sensor for ultraviolet and blue light that controlled the direction of rotation of
their flagella to avoid these dangerous rays. Since these flagella act as propulsion pro-
pellers, the cells could now not only see the harmful short-wave light but also avoid it.
A descendant of this blue light sensor is still found today in many primitive protozoa.
This ingenious blue light sensor probably also served the cells as a construction
manual for a solar collector, thanks to which they could feed on the energy of sunlight.
Cells shifted the blue sensor’s absorption to yellow-orange to capture as much of the
sunlight’s energy as possible. The cells coupled this solar collector to a system that con-
verted the captured light into chemical energy. With its help, the cells could now power
energy-hungry processes such as growth, cell division, and movement, or synthesizing
fat, sugar, and proteins. This primitive photosynthesis is still found today in some single-
celled organisms that thrive in the salt-rich margins of the Dead Sea or spoiled cured
fish. Ultimately, however, this form of photosynthesis proved to be a dead end because
it did not efficiently convert light from the sun into chemical energy. When later cells
used chlorophyll as a solar collector, ushering in modern photosynthesis, photosynthesis
that evolved from the blue light sensor remained limited to a few primitive single-celled
organisms.
So, life learned very early to see the world in two colors—blue and yellow-orange.
And now that it had seen color, it no longer wanted to do without it. With their extensive
and information-rich genetic material, the complex modern cells created three, four, or
even five different variants from the primitive blue light sensor, which opened up a vast
and differentiated color spectrum for them. Even more, these modern cells were able

[Link]
VIII � Foreword

to couple the signals of these different color sensors separately to increasingly complex
nervous systems.
Our eyes are equipped with five different light sensors, all closely related chemically
and probably descended from the primordial blue light sensor mentioned earlier. One of
these sensors is not used for vision but for the daily calibration of our “circadian” body
clock. Another sensor is found in the rod cells of our retina. This sensor is susceptible
to light and, therefore, we use it in dim light. However, this high light sensitivity comes
at a price because our retinal rods do not detect color or fine detail. In bright light, we
use three color sensors in the cone cells of our retina—one for blue, one for green, and
one for red. These sensors are not very sensitive to light, but they show us fine detail—
and color. Since each of these three color sensors can detect about a hundred different
intensities of color, and our brain compares the signals from the three sensors, we can
see not just three but one to two million colors. Older animals, such as insects and birds,
have up to five different color sensors and cannot only distinguish many more colors
than we humans can, but some can also see ultraviolet or infrared light to which we
are blind. When the first mammals evolved, they mainly hunted at night, leaving some
of their color sensors to atrophy, leaving only two of them. Almost all mammals—such
as dogs, horses, cats, and cows—therefore see only about 10000 different colors—about
the same as “color blind” humans. Only when intelligent apes wanted to distinguish
ripe from unripe fruit against the background of multicolored leaves did they again
develop a third color sensor, which allowed them to see the world in a new blaze of
color. So, humans and our close relatives, the great apes, are the only mammals that can
see millions of colors.
In this impressive book, Ingo Klöckl describes the magic realm of colors from a
chemist’s perspective. The synthesis of modern dyes with intoxicating color depth and
impressive stability was one of the great triumphs of nineteenth and twentieth century
chemistry, and the development of rewritable digital data carriers or catalysts for light-
driven water splitting suggests that the time of color chemistry is far from over. Ingo
Klöckl describes the bewildering variety of dyes available today and gives us detailed
information on how they can be produced, categorized, and compared with each other.
This book is a masterpiece, a true magnum opus that reveals to us in each chapter a new
wonder from the world of colors. The wealth of information it imparts to us is almost
mind-boggling, yet it is an exciting read for anyone who is no stranger to chemistry. The
book also builds a most welcome bridge between science and art, which have become
increasingly distant from each other in recent centuries, forgetting their common roots.
May not only natural scientists but also painters and art scholars pick up this book and
lose themselves in the magical realm of colors.

Basel Dr. Gottfried Schatz†


Foreword to the English edition
Dear readers, painting scientists, and researching painters, when the German edition
appeared, I never expected it would appeal to so many fellow human beings since the
balancing act between painterly observations and chemical-physical theories presup-
poses a profound understanding in many areas or at least interest. However, I was
proven wrong, so this English edition will hopefully accompany and support your work
in the vast, intricately interwoven field of art and technology.
Unfortunately, the unhelpful division into natural science and humanities also di-
vides the view of our world as the gods of the Greeks divided the spherical people. Buck-
low’s work [10] showed me how a holistic, Platonically oriented understanding of the
science of painting looked in the Middle Ages and embraced (al-)chemistry and painting
equally.
My deepest hope is that you will also see this bridging of art and science as a con-
tribution to an overall humanistic understanding of the world, as was familiar to many
great minds of science.

[Link]
Acknowledgment
My first and most important thanks go to my enchanting wife, Claudia, who once again
had to spend a large part of her time with a fanatically lecturing author and who,
through constant gentle urging, got me to finish the incubation of the book. Only this
made the publication possible at all. Moreover, most important, she dedicated vast
amounts of her time to translating the text into readable English, urging me to improve
and clarify numerous statements. She introduced me to the secrets of translation-
oriented writing and the benefits of a concise language.
Sadly, the author of the foreword, Dr. Schatz, passed away, to whom I will forever
be grateful for the insightful discussions and his foreword, which so well expressed the
magic of color. Furthermore, I would like to thank Mr. D. Widmer for important book
recommendations and information on writing inks, Dr. S. Hunklinger for an engaging
discussion on the color of semiconductors, Dr. T. Vilgis, Dr. B. Schneppe, and G. Bosse for
a discussion of the topic of egg white binders and clarea, Dr. W. Müller for his feedback
on the composition of acrylic paints, Dr. B. Born and Mrs. R. Ardal-Altun for a long, in-
formative and entertaining telephone conversation on the production of artists’ papers,
Dr. G. Kremer for many valuable suggestions, improvements and information from the
practice of a paint manufacturer, Dr. K.-O. Schäfer and Dr. W. Thiessen for information on
ink production, and Dr. G. Schatz for sending additional material. I would especially like
to thank Dr. Kremer, and again Dr. Schatz, for their positive evaluation of the manuscript.
They encouraged me, once as a practical color chemist and once as a versatile natural
scientist, to pursue the book’s aim. They were right, as the friendly and positive letters
from colleagues and experts prove, and I thank them for their interest and constructive
comments.
I would also like to thank the artists Mrs. S. Steinbacher and Mr. H. Karlhuber, with
whom I learned old-master and contemporary painting practice and who were thus
“question suppliers.”
Finally, I would like to thank Dr. R. Sengbusch, Mrs. Bentkuvienė, and the production
team of De Gruyter-Verlag as competent publication partners.

[Link]
Table of contents for volume 1
Foreword � VII

Foreword to the English edition � IX

Acknowledgment � XI

1 Introduction � 1
1.1 Further (and underlying) literature � 5
1.2 Pigments as the basis of painting � 6
1.2.1 Pigments of the ancient and early medieval world � 6
1.2.2 Pigments from the late Middle Ages onward � 10
1.2.3 Pigments from Romanticism and Classicism to
Classical Modernism � 12
1.2.4 Modern pigments � 14
1.3 Overview of pigments � 18
1.3.1 Color range white, black � 18
1.3.2 Color range yellow-orange-brown � 19
1.3.3 Color range red, purple � 21
1.3.4 Color range blue � 22
1.3.5 Color range green � 24
1.4 Paint systems, definitions � 27
1.5 Basic physical processes, spectra � 28
1.5.1 Emission colors � 29
1.5.2 Absorption color � 29
1.5.3 Color by absorption at a band edge � 33
1.5.4 RGB and CMY primaries, tristimulus theory, metamerism � 33
1.6 The interaction of light and matter � 37
1.6.1 Basics of dielectric materials � 37
1.6.2 Microscopic view: the oscillator model � 39
1.6.3 Macroscopic view: absorption � 44
1.6.4 Macroscopic view: Absorption by size-dependent collective excitations,
surface plasmons � 45
1.6.5 Macroscopic view: transmission, refraction, dispersion � 51
1.6.6 Macroscopic observation: scattering, reflection, brilliance � 55
1.6.7 Consequences of absorption: metallic luster, metallic colors,
bronzing � 62
1.6.8 Consequences of scattering: opacity, white pigments, and depth
light � 66
1.7 Summary: physical factors influencing pigment properties � 70
1.7.1 Particle size � 71
XIV � Table of contents for volume 1

1.7.2 Crystal structure and particle shape � 72

2 The chemistry of color � 77


2.1 Chemical absorption mechanisms � 77
2.2 SC: band gap transitions in semiconductors � 82
2.2.1 Valence and conduction band � 84
2.2.2 Color � 86
2.2.3 SC-based chromophores � 88
2.2.4 Influence of lattice width and crystal structure, thermochromism � 92
2.2.5 Alloys, solid solutions, and color � 95
2.2.6 Manufacture of semiconductor alloys � 100
2.2.7 Doping and blue diamonds � 101
2.3 LF: splitting d orbitals in a ligand field � 102
2.3.1 Crystal field theory and ligand field theory � 104
2.3.2 Splitting of degenerated d orbitals � 104
2.3.3 Spectroscopic selection rules � 111
2.3.4 Ligand-field splitting in octahedrally coordinated complexes � 113
2.3.5 Influence of ligand field strength � 121
2.3.6 Distortion of the octahedral field, Jahn–Teller effect � 125
2.3.7 Tetrahedral coordination � 127
2.3.8 LF-based chromophores � 130
2.4 CT: Charge transfer transitions � 131
2.4.1 Ligand-to-metal transition and oxygen-to-metal transition � 134
2.4.2 Metal-to-metal transition (MMCT), intervalence transition (IVCT) � 138
2.5 MO: molecular orbital transitions � 143
2.5.1 VB and MO model, resonance structures � 148
2.5.2 Chromophore enlargement, bathochromic shifts � 150
2.5.3 Donor–acceptor chromophores � 150
2.5.4 Polyene chromophore � 164
2.5.5 Polymethine chromophores � 183
2.5.6 Other chromophores: Sulfide radical ions � 189
2.6 Laking and colored lakes � 190
2.6.1 Structure of the color lakes � 194
2.6.2 Practical procedure � 197
2.6.3 Hue shift � 199

3 Inorganic pigments � 205


3.1 Carbon pigments � 207
3.2 Copper pigments � 217
3.3 Ultramarine pigments � 227
3.4 Oxide and sulfide pigments � 231
3.4.1 Classical heavy metal oxides and sulfides � 231
Table of contents for volume 1 � XV

3.4.2 Iron oxide pigments, ocher � 238


3.4.3 Complex inorganic color pigments (CICP), mixed metal oxides
(MMO) � 254
3.4.4 Cerium sulfide pigments � 266
3.4.5 Chromium oxide pigments � 268
3.4.6 Titanium oxides and zinc oxides � 270
3.4.7 Cadmium sulfide pigments � 274
3.5 Bismuth pigments � 278
3.6 Chromium pigments � 280
3.6.1 Chromate and molybdate pigments � 280
3.6.2 Chrome green (PG15, CI 77510); fast chrome green (PG48, CI 77600); zinc
green � 284
3.7 Iron blue pigments (Prussian Blue, Berlin Blue, Milori Blue, Paris Blue, iron
blue, PB27, CI 77510, 77520) � 285
3.8 Various metal pigments � 288
3.8.1 Calcium carbonates � 288
3.8.2 Lead white, flake white, Kremser white, Cremnitz white (PW1) � 289
3.8.3 White sulfates � 290
3.8.4 Miscellaneous colored pigments � 291
3.9 Glasses � 293
3.9.1 Glass coloring � 298
3.9.2 “Decolorization” of glass, color compensation � 302
3.9.3 Ancient glass coloring � 303
3.9.4 Frit colors � 305
3.9.5 Opaque glass � 307
3.10 Enamel � 307

4 Organic colorants � 309


4.1 Natural organic colorants � 310
4.2 Synthetic organic colorants � 312
4.2.1 Meaning of molecular structure � 315
4.3 Carotenoids � 317
4.3.1 Xanthophylls � 317
4.4 Flavanoids � 319
4.4.1 Origin in metabolism � 319
4.4.2 Classification � 320
4.4.3 Flavan-3-ols (catechins), flavan-3,4-diols, and flavanones � 321
4.4.4 Flavones � 322
4.4.5 Anthocyanins � 327
4.4.6 Neoflavones � 334
4.4.7 Quinone methides � 336
4.4.8 Chalcones and quinochalcones � 337
XVI � Table of contents for volume 1

4.4.9 Cause of color � 339


4.5 Xanthones � 341
4.6 Quinones � 343
4.6.1 Vat dyeing � 344
4.6.2 Natural quinones and naphthoquinones � 345
4.6.3 Natural anthraquinones � 346
4.6.4 Synthetic quinones � 353
4.6.5 Cause of color � 356
4.7 Indigoid colorants � 357
4.7.1 Natural indigoid colorants � 358
4.7.2 Synthetic indigoide colorants � 362
4.7.3 Dyeing with indigo and derivatives � 364
4.7.4 Cause of color � 364
4.8 Polymethine colorants: di- and triarylmethines, quinone imines � 365
4.8.1 Triarylmethine colorants � 367
4.8.2 Diphenylmethines, diarylmethines, indamine dyes � 378
4.9 Dioxazine pigments � 381
4.10 Phthalocyanine pigments � 382
4.11 Azo colorants (hydrazone colorants) � 387
4.11.1 Azo-hydrazone tautomerism � 389
4.11.2 The diazo component � 389
4.11.3 The coupling component � 390
4.11.4 Classification of azo pigments (hydrazone pigments) � 395
4.11.5 Cause of chromaticity, blue and green azo colorants � 412
4.11.6 Chronix toxicity, carcinogenicity � 417
4.12 Quinacridone pigments � 418
4.13 Perylene pigments � 421
4.14 Diketopyrrolo-pyrrole (DPP) pigments � 423
4.15 Azomethine, methine or isoindoline pigments � 424

Bibliography � 429

Index � 473
Table of contents for volume 2
Foreword � VII

Foreword to the English edition � IX

Acknowledgment � XI

5 Dyes for writing, painting, and drawing � 499


5.1 Types of bonds in the dye-substrate system � 501
5.2 Paper as a dye carrier � 503
5.2.1 Paper as ink carrier � 503
5.2.2 Paper dyeing � 504
5.2.3 Modification to paper compatible dyes � 506
5.3 Reactive dyes � 510
5.4 Direct or substantive dyes � 513
5.5 Mordant dyes (metal complex dyes) � 518
5.6 Cationic dyes � 522
5.7 Anionic or acid dyes � 524

6 Structure of paint systems � 529


6.1 Binders � 530
6.1.1 History � 530
6.1.2 Binder analytics � 534
6.1.3 Drying of binders � 536
6.1.4 Bonding types in the world of colorants � 538
6.2 Solvents � 541
6.3 Wetting agents and dispersants, grinding paints � 542
6.3.1 Wetting agents � 543
6.3.2 Dispersants � 550
6.3.3 Stabilization of dispersions, dispersants � 552
6.4 Thickener, rheology modifier � 556
6.5 Film-forming aids (coalescing agents) � 565
6.6 Other excipients � 566
6.7 Paper � 567
6.7.1 Structure and composition of raw materials � 567
6.7.2 Pulp from wood � 576
6.7.3 Composition and manufacture of paper � 589
6.7.4 Sizing and coating � 605
6.7.5 Calendering (satinage) � 613
6.7.6 Paper grades, general and industrial � 614
6.7.7 Special case artists’ paper � 616
XVIII � Table of contents for volume 2

6.7.8 Paper decay � 622


6.7.9 Aging-resistant paper � 624
6.7.10 Yellowing � 624

7 Paint systems in art � 627


7.1 Ceramics and their painting � 627
7.1.1 Classical ceramic painting � 627
7.1.2 Pigments of the cold-painting technique � 630
7.1.3 Ceramic enamel and glaze colors � 630
7.2 Stained glass � 640
7.2.1 Reverse glass painting � 642
7.2.2 Stained glass windows � 644
7.2.3 Stained glass � 646
7.3 Fresco (mural painting) � 649
7.3.1 Fresco-buono technique � 650
7.3.2 Lime-painting technique � 651
7.3.3 Fresco-secco technique � 651
7.3.4 Mixed techniques � 652
7.3.5 Pigment degradations � 652
7.4 Oil paint � 655
7.4.1 Basic composition of oil paints � 655
7.4.2 Types of oils � 656
7.4.3 Drying of oils, film formation � 660
7.4.4 Stand oils � 672
7.4.5 Effect of heavy metals, siccatives � 675
7.4.6 Linseed oil varnish � 676
7.4.7 Technical improvement of colorants and painting agents in the nineteenth
century, paint tubes � 676
7.4.8 Resins, resin balsam, turpentine oil � 679
7.4.9 Other solvents: benzines, turpentine substitutes � 687
7.4.10 Varnish materials � 689
7.4.11 Pigment degradations � 691
7.5 Protein systems (poster paint, gouache paint, glue paint, size paint,
distemper paint) � 703
7.5.1 Albumin as binder (whole egg, egg white) � 706
7.5.2 Collagen as a binder (poster, gouache, glue, size, distemper paint) � 709
7.5.3 Casein as binder � 713
7.6 Tempera � 717
7.6.1 Egg yolk tempera, pure egg tempera � 718
7.6.2 Egg tempera � 720
7.6.3 Fatty egg tempera, egg-oil emulsions � 720
7.7 Watercolors � 721
Table of contents for volume 2 � XIX

7.7.1 Basic composition of watercolors � 721


7.7.2 Gum Arabic � 723
7.7.3 Gum tragacanth � 725
7.7.4 Ox gall � 726
7.7.5 Paper � 726
7.8 Alkyd colors � 727
7.9 Acrylic paints � 731
7.9.1 Basic composition � 732
7.9.2 Irreversible film formation � 734
7.9.3 Retarders � 738
7.9.4 Media, thickeners, gels, acrylic butter � 738
7.9.5 Wetting agents and dispersants � 739
7.9.6 Film formation aids � 740
7.9.7 Other additives � 741
7.10 Lithographic printing, lithography � 741
7.10.1 Lithographic crayons and inks � 744
7.10.2 Reprint and materials for reprint � 744
7.10.3 Printing inks for lithography � 745
7.11 Silicate paint � 745
7.12 Low binder systems: chalks and pencils � 746
7.12.1 Blackboard chalk � 746
7.12.2 Pastel crayons � 747
7.12.3 Pencils � 747
7.12.4 Colored pencils � 750
7.12.5 Paper � 750
7.13 Fingerpaint � 750
7.14 Intarsia art � 751

8 Inks � 753
8.1 Carbon inks � 755
8.1.1 Inks in antiquity � 756
8.1.2 Modern carbon inks � 760
8.1.3 Chemistry of carbonization, combustion, and sooting � 761
8.2 Chemistry of phenolic ink constituents � 775
8.2.1 Oxidation and polyphenols � 776
8.2.2 Hydrolyzable tannins � 779
8.2.3 Condensed or nonhydrolyzable tannins, proanthocyanidins � 781
8.2.4 Tannin-like tanning agents � 783
8.3 Inks based on natural materials, book illumination � 787
8.3.1 Colored natural inks, book illumination � 787
8.3.2 Brown inks � 792
8.4 Durable writing inks (iron gall inks) � 793
XX � Table of contents for volume 2

8.4.1 Chemistry of iron gall inks � 796


8.4.2 Color of iron gall inks � 797
8.4.3 Brown iron inks � 798
8.4.4 Excursion: the iron-phenol reaction � 800
8.5 Dye inks (fountain pen, felt-tip pen, ballpoint pen, inkjet printing) � 801
8.5.1 Function of components � 806
8.5.2 Colorants for fountain pen ink � 815
8.5.3 Colorants for felt-tip, fiber-tip, ballpoint pens � 818
8.5.4 Colorants for inkjet inks � 818
8.5.5 Paper, inkjet support materials � 826
8.5.6 Colorants for stamp pads � 832
8.6 Laser or copier toner � 832
8.7 Printing inks � 836
8.7.1 Inks � 841
8.7.2 Pigments � 843
8.7.3 Binders � 846
8.7.4 Rosin derivatives as binders � 849
8.7.5 Solvents � 852
8.7.6 Auxiliaries � 854
8.7.7 Paper � 855
8.8 Tusche � 855
8.8.1 Sepia, Natural Brown 9 � 857
8.8.2 Shellac � 858

Bibliography � 861

Index � 903
5 Dyes for writing, painting, and drawing
In the subject of “dyes,” we risk entering a varied and wide-ranging field that can quickly
lead us away from the book’s actual subject. Nevertheless, we want to focus on some
basic properties of dyes since certain forms of expression in the visual arts, for exam-
ple, printmaking and calligraphy, use pigments and dyes alike to achieve visible results.
Later in this book, we will highlight printing inks that are pigmented or composed of dye
pastes, (book) painting that works with (dye) inks and ballpoint pens, and the adhesion
of watercolor colorants to paper or fabric. Albeit inadequately, we will address the often
complex mixtures of natural dyes used in book illumination in ▶Section 8.3.
Unfortunately, we have to limit ourselves to a selection of well-defined, more mod-
ern colorants relevant to artists’ materials, and we will not consider the history and
chemistry of dyeing. This exciting subject is addressed by [13, 18–21, 109]. These works
provide good introductions for readers interested in textile dyeing. In addition, encyclo-
pedias such as [203] contain dedicated sections on this field of application.
Due to its long history and huge economic significance, a true plethora of dyes are
available today. To discuss them, we can classify them according to different character-
istics, which cannot be strictly distinguished from each other and partly overlap. The
following categories are relevant for this topic:
– solubility (for arts, mainly in water, vegetable oils, turpentine oil, and solvents based
on mineral oil)
– dying technique
– chemical structure

Further classifications are possible and in use depending on the context, such as chem-
ical structural types or dyes for special applications (leather, hair, stationery).
Since water plays a significant role as a solvent in the arts, we first organize col-
orants according to their water solubility, progressing to the dying technique, and finally,
to their chemical structure, ▶Figure 5.1:
– Solvent dyes. They are colorants insoluble in water but soluble in other solvents. This
property is caused by the absence of polar groups such as sulfonic acids, carboxylic
acids, or ionic centers. Therefore, they are suitable for coloring plastics, oils, and
waxes, and in the field of artists’ materials, for filling pencils and pens.
– Disperse dyes. They are sparingly soluble in water and employed molecularly dis-
persed. First, they are adsorbed superficially on the substrate and then diffuse into
it. However, their field of application is not the art sector but the dyeing of hydropho-
bic fibers made of polyester, nylon, or acrylic.
– Development dyes. The color body forms only during dyeing from mostly colorless
precursors on the substrate or near it. For example, vat dyes form from soluble col-
orless leuco bases, and mordant dyes from soluble complex-forming dyes. The re-
sulting colorants are insoluble in water and applicable as artists’ pigments.

[Link]
500 � 5 Dyes for writing, painting, and drawing

Figure 5.1: Classification of dyes according to their solubility in water, dyeing technique, and chemical
structure.

– Reactive dyes. Their main feature is forming covalent bonds with the substrate dur-
ing dyeing, thus acquiring high wash fastness. In order to achieve this, the colored
structure contains reactive groups adapted to the substrate, such as cotton, wool,
nylon, and paper.
– Anionic or acid dyes are water-soluble anionic compounds mainly used for nylon,
wool, silk, acrylic fibers, paper, leather, food, and inks. They are addressed as direct
dyes when they have a high affinity for cellulosic substrates.
– Direct dyes. Their main characteristic is “substantivity,” i. e., strong adsorption to cel-
lulose and cellulose derivatives (cotton and paper) from an aqueous solution with-
out further auxiliaries. They are anionic water-soluble dyes, and thus a subgroup
of acid dyes.
– Cationic or basic dyes. They consist of water-soluble cationic compounds and are
mainly suitable for paper, polyacrylonitrile, nylon, and polyester.

▶Table 5.1 summarizes the substrate’s adhesion mechanism for the indicated dye
classes. The English names of the classes correspond to a standard naming scheme
for dyes according to the Color Index [1], e. g.,
– RR24 (Reactive Red 24) denotes a red reactive dye for inks.
– DBk19 (Direct Black 19) denotes a black direct dye for inkjet printing.
– BG4 (Basic Green 4) or AY23 (Acid Yellow 23) indicates green and yellow dyes for
felt-tip pens and inks.
5.1 Types of bonds in the dye-substrate system � 501

Table 5.1: Different dye classes and mechanisms to achieve the adhesion to the substrate [19, 20].

Class Mechanism of adhesion

Disperse dye Dipole-dipole interaction, dispersion forces


Vat dye Dipole-dipole interaction, dispersion forces
Mordant dye Aggregation, dipole-dipole interaction, dispersion forces (ionic bonding)
Reactive dye Covalent bonding
Acid dye (anionic dye) Ionic bonding, dipole-dipole interaction, dispersion forces
Direct dye Ionic bonding, aggregation, dipole-dipole interaction, dispersion forces
Basic dye (cationic dye) Ionic bonding, aggregation

5.1 Types of bonds in the dye-substrate system


Before we look at individual dye classes, we first briefly summarize the mechanisms
that provide adhesion of the dye to the substrate. They are discussed in more detail in
▶Section 6.1.4. In dye chemistry, we encounter mainly noncovalent bonding modes and
physical intercalation of larger aggregations in fiber voids, ▶Table 5.1.

Ionic bonding
Apart from reactive dyes, where covalent bonding is decisive, ionic bonding plays a cru-
cial role in the dyeing process. Examples are anionic acid dyes, forming ionic bonds
with ammonium cations of proteins. Conversely, cationic dyes interact with carboxy-
late anions of protein fibers. Typical ionic sites of common fiber materials are listed
in ▶Table 5.2. We see that protein fibers and some synthetic fibers naturally provide
binding sites. At the same time, other materials such as high-quality paper tend to react
nonionically and must be coated to improve adhesion. ▶Section 6.7.4 shows how this is
accomplished with paper.

Table 5.2: Typical ionic sites of common fiber materials, relevant for dyeings [19, 20].

Fiber Cationic sites Anionic sites

Wool Proteinic amino groups Proteinic carboxyl groups


Silk Proteinic amino groups Proteinic carboxyl groups
Nylon Terminal amino groups Terminal carboxyl groups
Cotton None None (hydroxyl groups not acidic)
Paper containing wood None Lignic acids and phenolates, glucuronic acids
Paper wood-free None Glucuronic acid contents
– glue-coated (Proteinic amino groups) (Proteinic carboxyl groups)
– polymer-coated Depending on polymer Depending on polymer
– rosin acid metal salt-coated Cationic metal salts None
502 � 5 Dyes for writing, painting, and drawing

Dipole-dipole interaction
A common factor in the dyeing process is the interaction of two permanent dipoles
formed from covalently bonded atoms of different electronegativity. The electrons in-
volved are permanently unequally distributed. Typical partial structures in dyes are
−OH, −CO−, −NH−, or −CN.
The hydrogen bond, a dipole-dipole interaction between hydrogen as a positive
partner and electronegative elements (O, N), has approximately tenfold strength. It oc-
curs in most groups of substances relevant for coloration.

Dispersion forces, actual van der Waals forces


All polymeric substrates (fibers of protein, nylon, and cellulose, and polyethene films)
have nonpolar and uncharged regions that can interact with dyes via temporary dipoles.
Therefore, extended nonpolar regions are created in a dye to mirror the substrate and
attach the dye to it. As a rule, such molecule parts are elongated or planar in structure.
We can observe the strength of this interaction in crystalline regions of polyethene,
which has no special features such as polar centers or charges. Nevertheless, these crys-
tallites show a high strength.

Hydrophobic interaction
Water shows a higher-order state at hydrophobic surfaces, as it aligns itself to the surface
and forms hydrogen bonds with itself. When the hydrophobic groups come together,
structured water molecules are released, and a lower-order state is reached. The entropy
gain is the driving force of the process. Hydrophobic interactions support apolar dye
regions’ adhesion to apolar substrates or the aggregation of dye molecules, like van der
Waals interactions.

Aggregation
The formation of aggregates is not an actual type of chemical bonding, but it describes a
vital adhesion mechanism summarily. It is based on the fact that dye molecules diffuse
into the fiber cavities and form larger aggregates. Driving forces can be all of the pre-
viously mentioned interactions. Subsequently, the complexes are physically trapped in
the fiber cavities. Aggregation is achieved in two ways:
– The first method is to cool down a hot dye bath. The dye is structured to show a
natural tendency to aggregate. Direct dyes have, e. g., extended nonpolar regions
that exert van der Waals forces against neighbor molecules. Aggregation is the nor-
mal state at a lower temperature and is reduced by heating the dye solution for
the dyeing process. After diffusion of the isolated dye molecules into the fiber, they
aggregate upon cooling and are physically retained in the substrate.
– The second method is forming a metal complex. Dye molecules form complexes with
metal salts after diffusion into the fiber. By design, the complex comprises two or
5.2 Paper as a dye carrier � 503

three dye molecules and is better retained in the fiber than a single dye molecule
since all of the weak interactions mentioned above now act on it many times their
strength. Furthermore, the complex is physically trapped due to its sheer size.

5.2 Paper as a dye carrier


This chapter presents information about dyes, which essentially comes from the field
of (textile) dyeing. In the beginning, textile dyes were also used to a large extent for the
dyeing of paper (in the mass, on the surface, or as a drawing), and numerous findings
originate from the commercial dyeing industry. Later, we will see how dyes specifically
developed today for paper or inkjet printing fit into the existing classification.

5.2.1 Paper as ink carrier

We will find the basic structure of the paper explained in detail in ▶Section 6.7. For now,
a first glimpse shows us that there are two main types of paper:
– Unbleached paper contains wood and has a high proportion of lignin in addition to
the desirable cellulose and hemicellulose. Besides, this polymer of phenylpropane
units has a high number of anionic substituents due to the preparation of the wood.
– Bleached wood-free paper consists ideally of pure cellulose and hemicellulose.

The chemistry of paper dyeing must therefore deal primarily with the properties of cel-
lulose. It is significant that, unlike natural textile fibers such as wool, or synthetic fibers
such as nylon, cellulose has no amino or amide groups, and thus no cationic centers.
In addition, its hydroxyl groups do not react acidic, so we also would not expect an-
ionic centers. However, as shown in ▶Section 6.7.3, a small portion of glucose is present
as glucuronic acid, and acidic hemicelluloses remaining in the paper pulp contribute a
more significant proportion of mannuronic and galacturonic acids to the paper’s anion
activity.
There are three stages to consider for dyeing:
– First, the dye diffuses into the cavities of the paper fibers.
– Second, the dye is adsorbed there.
– Finally, the dye’s insolubility is increased to improve smear and water resistance by
staining with metal salts and adding fixatives (▶p. 602) or binders.

The coloring of the paper itself is usually not the artist’s responsibility, as the processing
happens in the paper mill. The dye can be added to the pulp mass before making sheets
or brushed or dipped onto the paper surface. Surface dyeing is better done with pig-
ments, reducing the risk of bleeding. With this type of dyeing, the paper manufacturer
can take measures to ensure good adhesion of the dyes on or in the paper pulp.
Conversely, when drawing or writing on paper, the artist cannot influence the effec-
tiveness of the three steps. In contrast to the textile dyer, neither can he choose reaction
504 � 5 Dyes for writing, painting, and drawing

conditions that lead to good adhesion of the dye, nor can he use hot dye baths, and thus
different solubilities or aggregation tendencies of the dyes. It is also impossible for him to
create the necessary acidic or alkaline environment, add additives, or hotfix the dyeing.
Therefore, it should come as no surprise that writing inks, in particular, are often
not very water-resistant and smear once they get wet. Possibilities for improving the
situation are:
– Papers are coated with a high affinity to dyes. We will become familiar with coating
materials, ▶Section 6.7.4.
– Binders added to the ink increase smudge and water resistance by forming a pro-
tective film. Discussing the composition of inks, we will get acquainted with appro-
priate means, ▶Section 8.5.
– Dyes are used to match the specific writing and drawing situation. We will see an
example of a modified direct dye in the following. Due to high development costs,
dyes are modified only for profitable markets such as inkjet or laser printing, not
for artists’ inks.

5.2.2 Paper dyeing

The mentioned standard works on dyeing always contain supplements on paper dyeing;
more details can be found in [203, keyword “paper”], [13, Chapter 5.3], [178, Chapter 9.3],
[177, Chapter 6.4], [179, Chapter 3.6.1], [1005], [205, p. 446ff]. In ▶Table 5.3, we see the
main dye classes considered for use in paper, namely anionic and cationic direct dyes
as well as basic dyes. ▶Table 5.4 contains examples of colorants currently offered for
paper coloring.

Table 5.3: Suitability of dye classes for paper dyeing [177, Chapter 6.4].

Characteristic Cationic or basic Anionic or acid dyes Direct dyes Direct dyes
dyes anionic cationic

Molecule size Small Small Large Large


Charge Positive Negative Negative Positive
Affinity toward
mechanical pulp + −2 − o
pulp unbleached o −2 o o
pulp bleached − −2 + +
Lightfastness − − o o
Market share1 28 % 2% 50 % 6%
Application Wood-containing Wood-containing and Wood-free papers Unsized or
papers, packaging wood-free sized alkaline
papers writing papers sized papers
1
[179, Chapter 3.6.1], the rest: pigments.
2
Requires cationic fixative, ▶p. 602.
5.2 Paper as a dye carrier � 505

Table 5.4: Example overview of pigments and dyes for paper coloring (surface and bulk) [981, 983, 984].

Colorants
Yellow AY1, AY17, AY23, AY36, AY42, BY2, BY28, BY29, BY37, BY40, BY51, BY57, BY87, BY96, DY11, DY44,
DY118, DY147
Orange AO7, AO10, AO51, AO56, BO1, BO2, BO31, DO15, DO26, DO102
Red AR1, AR14, AR18, AR52, AR73, AR88, AR97, AR119, AR131, AR151, BR12, BR14, BR15, BR18, BR46,
DR6, DR16, DR23, DR80, DR81, DR236, DR239, DR254
Purple AV12, AV17, AY49, BV1, BV2, BV3, BV4, BV5, BV8, BV10, BV11, BV14, BV16, DV9, DV35
Green AG1, AG20, BG1, BG4, DG26
Blue AB1, AB7, AB9, AB92, AB113, BB1, BB3, BB4, BB7, BB9, BB26, BB41, BB99, BB159, DB15, DB71,
DB80, DB86, DB199, DB218, DB290
Black ABk1, ABk2, ABk52, ABk172, ABk210, ABk234, DBk19, DBk155, DBk170, DBk186
Brown BBr1, BBr4, BBr16, BBr17
Metal complexes for muted colors
Yellow AY59, AY99, AY114, AY194, AY204, AY241
Orange AO74, AO86, AO142
Red AR183, Ar184, AR186, AR194, AR195, AR219, AR357, AR362
Purple AV90, AV92
Green AG104
Blue AB158, AB193
Black ABk2, ABk52, ABk58, ABk60, ABk194, ABk172
Brown ABr45, ABr355, ABr365, ABr369
Pigments
Yellow PY1, PY3, PY12, PY13, PY14, PY17, PY65, PY74, PY83, PY174, PY191:1
Orange PO5, PO13, PO16, PO34
Red PR2, PR3, PR8, PR19, PR22, PR23, PR48:1, PR48:2, PR48:3, PR49:1, PR49:2, PR53, PR53:1, PR57:1,
PR60:1, PR63:1, PR112, PR146, PR170
Purple PV19, PV23
Green PG7, PG8
Blue PB15, PB15:1, PB15:3, PB15:4, PB29
Black PBk7

These classes are selected from the viewpoint of a paper manufacturer, for whom a high
coloring efficiency is essential. In writing or drawing fluids, in contrast, it may well make
sense to employ, e. g., acid dyes. For this application, poor adhesion, manifesting as a
lack of smear or water resistance, can be tolerated or improved by adding binders to
the writing fluid.
Cationic dyes (basic dyes) are particularly suitable for wood-containing or un-
bleached wood-free papers since lignin and acidic hemicelluloses in the aqueous paper
pulp have a high anion activity during production. The anions originate from uronic
acids and artifacts due to preparation, such as phenolic hydroxyl groups and sulfonic
acids [203, keyword “cationic dyes”]. The dyes can adhere well due to ionic bonds.
However, their affinity to bleached paper is low, and anionic fixatives are needed here,
▶p. 602. Since the dyes are brilliant but not very lightfast, they are more suitable for
506 � 5 Dyes for writing, painting, and drawing

low-quality wood-containing papers, wrapping paper, and packaging material. Exam-


ples are BY2, BO2, BR1, BV1, BV2, BV3, BV4, BV10, BB9, BB11, BB26, BG1, BG4, and BBr1.
In contrast, anionic direct dyes show a high affinity to bleached paper. Their adhe-
sion via hydrogen bonds and van der Waals forces is good. Cationic fixatives can support
rich and deep colors if required in individual cases, ▶p. 602. They are suitable for col-
oring wood-free papers, such as writing papers, blotting papers, and napkins. Examples
are DO102, DR239, and DB218.
Cationic direct dyes are listed in the CI systematics under the cationic dyes but differ
from them in their size and linear or planar molecular shape. They adhere via hydrogen
bonds and van der Waals forces. The intentional introduction of cationic centers such
as −C2 H4 N⊕ (CH3 )3 further increases substantivity by forming ionic bonds. Cationic di-
rect dyes have a moderate affinity for bleached and wood-containing paper but a high
affinity for wood-free paper with its anion activity. Therefore, they are applicable for
unsized and alkaline-sized papers.
Less used (commercially) are acid dyes since they show little affinity to cellulose
and other paper pulps. Although they diffuse well into the cavities of the paper felt, they
always require the use of a cationic fixative (▶p. 602), resin sizing, or a metal salt (alu-
minum) that can precipitate color lakes. They are used for wood-containing and wood-
free, sized writing papers; examples are AO7 and AV17 (cationized).

5.2.3 Modification to paper compatible dyes

The special requirements for paper dyeing are as follows:


– high substantivity for the substrate cellulose
– short dyeing time at low temperature, and thus good cold water solubility
– improved lightfastness

Developments related to cold water solubility and increased substantivity, especially of


anionic direct dyes, follow several directions [13, Chapter 5.3] depending on the intended
application:
– Replacing a benzoyl group with the triazine ring and introducing additional hy-
droxyl groups via the bis-(2-hydroxyethyl)amino group. In the common dye DR81,
this modification leads to the increased solubility and substantivity of DR253, ▶Fig-
ure 5.2.
– Recharging, i. e., the introduction of positive charges. The substantivity toward an-
ionic paper pulps can be increased in the manner of cationic direct dyes. The N,N-
diethylaminopropyl side chain is responsible for introducing cationic charges; its
quaternary amine is counterbalanced as an acetate or lactate.
From DR81, we thus arrive at BR111, ▶Figure 5.2, in which these side chains are cou-
pled to the chromophore with cyanuric chloride. Phthalocyanines can also be made
paper compatible in this way. The N,N-diethylaminopropyl side chain is introduced
5.2 Paper as a dye carrier � 507

Figure 5.2: Possibilities of modifying a direct dye to increase its suitability for paper dyeing: Intro-
duction of hydroxyl groups to form hydrogen bonds (DR253) and cationic groups to form ionic bonds
(BR111) [178, p. 138].

by treating the phthalocyanine with chlorosulfonic acid and amidating it as a sul-


fonamide:

– The use of the coupling component 2-amino-5-naphthol-7-sulfonic acid and its


N-acetyl, N-benzoyl, or N-aryl derivatives for soluble and substantive azo dyes.
Examples include DO102, DR239, DV51, or DB71:
508 � 5 Dyes for writing, painting, and drawing

– Employment of high molecular weight polykis azo dyes with the corresponding sub-
stantivity; examples are again DR81, DR253, and BR111 or DB71.
– The use of heterocycles as azo and coupling components in azo dyes such as DY28,
DY137, and DY147:
5.2 Paper as a dye carrier � 509

– Bridging mono and diazo dyes with phosgene or cyanuric chloride to obtain large
molecules with increased substantivity. Examples again include DO102 and DR239
as well as DY51 and DG26, in which a yellow and a blue dye have been combined to
a green preparation.

Dyes with higher lightfastness are obtained mainly by metallization with copper. The
copper complexes preserve the planar shape of the molecules, and thus the substantiv-
ity. 3,3’-Dimethoxy-4,4’-biphenyldiamine (dianisidine) has proven itself as a complex-
forming bridge, which forms metal complexes by demethylation. DB218 provides an
example:

Further developments replace toxic components such as the dianisidine bridge (a ben-
zidine derivative) with nontoxic bridges, such as, e. g., in DB273:
510 � 5 Dyes for writing, painting, and drawing

The designers of inkjet inks are confronted with similar problems; see ▶Section 8.5.4 for
the solutions adopted.

5.3 Reactive dyes


Reactive dyes [19, 20], [203, keyword “reactive dyes”], [13] are water-soluble dyes that
form a covalent bond with the substrate during dyeing, especially with the groups −OH,
−SH, or −NH−, as they are present in many textile fibers made of cotton, wool, or ny-
lon. Therefore, reactive dyes exhibit very high adhesion and washing fastness. They are
mainly applied in textile dyeing but also in writing and inkjet inks.
First experiments with this type of dye were carried out in 1895, and the first in-
dustrially important dye Supramin Orange R for wool came onto the market around
1938. It carries the reactive chloroacetylamino side chain −NH−CO−CH2 −Cl. From the
1960s onward, dye families with triazines, vinylsulfonic acids, and similar compounds
as anchors followed in rapid succession.

Chemistry
According to their mode of operation, reactive dyes consist of two parts: a chromophore
and a reactive group. During dyeing, this group couples with the fiber. The dye can carry
other groups that increase the solubility of the precursor in the aqueous dye bath:

A sulfonic acid −SO3 H is often applied as a solubilizing group. Any organic chromophore
can act as the color-active part of the dye. Often azo, anthraquinone, or phthalocyanine
chromophores are used so that reactive dyes can achieve complete coverage of all hues.
In contrast to other classes of colorants, large nonpolar or ionic molecular skeletons are
not required since covalent bonding ensures adhesion adequately.

Adhesion
The reactive group that ensures an anchoring of the dye to the fiber is multiform and
leads to different dye families. Frequently, triazines and pyrimidines or vinyl sulfonic
acids are used, which can be halogenated in different ways:
5.3 Reactive dyes � 511

The triazine anchor, e. g., is coupled to the dye via cyanuric chloride, ▶Figure 5.3. The
selective reaction of the three chlorine atoms (the first reacts at 0–5 ℃, the second at
35–40 ℃, the third at 80–85 ℃) allows controlled introduction of the dye. During the dye-
ing process, the anchor reacts with hydroxyl or amino groups of the dyeing material by
eliminating hydrogen chloride, forming an ether or amine bond. A similar reaction for
cellulose (paper) would require alkaline conditions to convert the hydroxyl groups of
the cellulose into anions.

Figure 5.3: Synthesis of a reactive dye with triazine anchor and coupling to the amino group of a fiber
[203, keyword “reactive dyes”], [19].
512 � 5 Dyes for writing, painting, and drawing

Examples
Of the following examples of reactive dyes with triazine anchors, RR24 is used for red
writing inks.

Anchors of vinylsulfonic acids form a carbon-carbon bond in an addition reaction, ▶Fig-


ure 5.4. For example, RR23 (red) is used in writing inks, RR180 (red) and RBk31 (black) in
inkjet inks.

Figure 5.4: Synthesis of a vinyl sulfone-based reactive dye and coupling to the hydroxyl group of a fiber
[203, keyword “reactive dyes”], [19].
5.4 Direct or substantive dyes � 513

5.4 Direct or substantive dyes


Direct or substantive dyes are water-soluble anionic dyes whose main characteristic is
substantivity, which means a very high dye affinity to natural or processed cellulose
(cotton, jute, viscose, paper) [13, 19, 20]. Direct dyes are among the oldest colorants. For
example, since antiquity, carotenoids such as crocetin and bixin have dyed textiles yel-
low (▶Section 4.3.1), whereas carthamine from the safflower made red dyeings possible,
▶Section 4.4.8 [609].
Direct dyes require for the dyeing no further auxiliaries such as mordants, sim-
plifying their use. The process runs in a neutral dye bath with NaCl or Na2 SO4 added.
A disadvantage is poor washing fastness, which a post-treatment of the dyeing material
improves; however, this does not quite meet today’s standards.
Of interest to us is their usage as dyes for writing and inkjet inks on paper. Unlike
other dye classes, direct dyes have a high affinity for lignin-free (bleached or wood-free)
paper, so paper production and dyeing require no extra fixative. They are employed
514 � 5 Dyes for writing, painting, and drawing

for coloring wood-free papers of all kinds, such as writing papers, blotting papers, and
napkins.
In paper chemistry, direct dyes are also called cationic direct dyes. According to the
CI system, they are classified as cationic dyes (basic dyes). However, they differ from
basic dyes in their size and linear or planar molecular shape and are thus closer to the
direct dyes. Therefore, they are suitable for dyeing wood-containing paper in contrast
to the anionic dyes.

Chemistry
The characteristic feature of direct dyes is a large, elongated, nonpolar molecule whose
solubility in water is increased by anionic groups, e. g., sulfonic acids. They share the
acid groups with acid dyes; in fact, they form a subgroup of the acid dyes. The transition
between high molecular weight acid dyes and direct dyes is fluent. However, unlike acid
dyes, they exhibit an extensive planar molecular shape, which provides the necessary
substantivity.
In terms of color range, direct dyes cover the entire spectrum. The examples from
antiquity show that polyene dyes can serve as a base. The advent of the azo chro-
mophore in modern times has given rise to a generation of direct dyes such as Congo
red. They have undergone further improvements like the azo chromophore itself. To-
day, azo dyes are used for the most part; for the dark, blue, brown, and black shades, bis
and polykisazo compounds and copper-azo complexes are employed. Copper phthalo-
cyanines provide greenish-turquoise shades.

Adhesion
The adhesion mechanism was previously thought to be a formation of an ionic bond
by the acid groups of the dye with the cationic centers of the substrate. Consequently,
hydrogen bonds between polar groups of the dye and the substrate would contribute
to adhesion due to the parallel orientation of dye and substrate. Hydroxyl groups of a
cellulose chain show an example:

Some factors suggest that these are not the only causes of adhesion: Cellulosic substrates
do not have the cationic positions at all, many vat dyes for cotton are not elongated but
only planar, and hydroxyl groups of the cellulose form hydrogen bonds with each other
5.4 Direct or substantive dyes � 515

rather than with a partner to which they are spatially insufficiently aligned [19, Chap-
ter 14.7]. Since an extended, planar molecular shape plays a significant role in the degree
of substantivity, today it is assumed that an interplay of the mentioned mechanisms with
van der Waals forces and aggregation occurs [19, Chapter 14.7], [20, p. 87].
The high affinity of anionic direct dyes to bleached paper is mediated by hydrogen
bonds and van Waals forces, ▶Section 5.2. We see below how incorporating the struc-
tural features from ▶Section 5.2.3 further improves the suitability for paper. Likewise, a
cationic fixative (▶p. 602) can improve rich and deep colors in individual cases if needed.
Cationic direct dyes adhere via hydrogen bonds and van der Waals forces; the cationic
centers increase substantivity by forming ionic bonds. Since lignin specifically has the
necessary anion charge, they are particularly suitable for use with wood-containing pa-
per.
Van der Waals and dispersion forces are only sufficiently strong for large dye
molecules. For this reason, size is of decisive importance for substantivity and can
serve as a criterion for differentiation from acid dyes. In the dye solution, hydropho-
bic interaction causes dye molecules to orient themselves parallel to molecular chains
of the fiber so that forces can act between nonpolar regions of dye and substrate.
It is assumed that cellulose, polar in nature, can also participate since its axial hy-
droxyl groups are more hydrophobic than equatorial hydroxyl groups or solvent wa-
ter.
The idea of aggregation is that individual dye molecules diffuse into the fiber cav-
ities and are adsorbed there. Then, driven by the hydrophobic interaction and favored
by their planar molecular shapes, they aggregate into larger units retained in the cavity.
Since aggregation of the individual molecules at room temperature also occurs in the
dye solution, the temperature must be increased during textile dyeing to dissolve the
aggregates and allow diffusion into the substrate.
The durability of the dyeing on textiles is less with direct dyes than with chemically
bonded dyes, but a suitable finishing achieves an improvement. Such treatment can
be performed with metal salts (preferably copper and chromium), whereby metal-azo
complexes are formed with hydroxy groups of the dye [19, Chapter 14.5]. The increased
molecular weight and decreased charge reduce the solubility and diffusivity of the com-
plex significantly. The hue shift toward darker tones (▶Section 5.5) is a side effect, which
is quite desirable for blue, brown, and black. The second possibility of post-treatment is
to diazotize free amino groups of the primary direct dye. A new azo dye is synthesized
directly in the substrate, which is also less diffusible and has a darker hue by coupling
with a suitable coupling component.

Examples
An early representative of the direct dyes is Congo red, which is no longer in use due to
its carcinogenic benzidine component:
516 � 5 Dyes for writing, painting, and drawing

Modern azo-based direct dyes are DBk19 and DR75, two dyes for writing inks. DBk19,
DBk154, and DBk168, as well as DY86 and DY132, are also used in black and yellow inkjet
inks:
5.4 Direct or substantive dyes � 517

Important examples of direct-acting phthalocyanine dyes are DB86 and DB199, used in
blue writing and inkjet inks:

The already excellent suitability of direct dyes for dyeing wood-free paper can be further
increased by modifying their structure, ▶Section 5.2.3. For example, the azo dye DB218
is a large molecule of high substantivity, and it is metalized to increase lightfastness.
DO102 and DR239 are built on 2-amino-5-naphthol-7-sulfonic acid as a coupling compo-
nent, increasing the dye’s substantivity and cold water solubility in paper applications.
In addition, the dye molecules are enlarged by linking two partial molecules with phos-
gene via a urea bridge, which also increases their substantivity:
518 � 5 Dyes for writing, painting, and drawing

5.5 Mordant dyes (metal complex dyes)


Mordant dyes often consist of an acid dye, which forms with the cations of a metal salt,
the so-called stain, a stable, insoluble complex, or color lake with high light and wash-
ing fastnesses [13, 19, 20, 46, 705, 706]. The laking that occurs in the process changes the
color, possibly significantly, ▶Section 2.6. Staining dyes were already used for dyeing
in antiquity. Classic examples are madder, carmine, and kermes (anthraquinone dyes),
dyer’s woad and dyer’s broom (flavonoid dyes), redwoods, and bluewoods (iso and ne-
oflavonoid dyes). All were laked with alum as an AlIII salt [609]. One specific dyeing with
madder yielded Turkish red, which was highly valued, ▶Section 4.6.3.
In modern dyeing chemistry, mordants significantly increase acid dyes’ light and
washing fastnesses by anchoring the dye in the substrate through complexation. Also,
the color shift opens up the range of blue, green, brown, and black.
Mordant dyes are not applicable for the production of artists’ material. They are
nevertheless interesting for us since, in past times, color lakes were obtained as pig-
ments from natural dyes, ▶Section 2.6.

Chemistry
Examples from antiquity show that a wide variety of structures inducing color can
serve as the basis for mordant dyes. Modern representatives are mainly based on an-
5.5 Mordant dyes (metal complex dyes) � 519

thraquinone (red to purple) or azo compounds (green, blue, brown, black). Due to the
metallization, the colors obtained are duller and darker than those of the original dyes,
▶Section 2.6.3, but also more lightfast.
For the formation of the metal complex, the addition of metal salts is required, es-
pecially CrIII , SnIV , AlIII , CuII , or FeIII salts. The addition aims at
– fixing the dye
– changing or deepening the hue

Both objectives are achieved by forming an insoluble metal-dye complex (color lake),
whereby the bathochromic effect described occurs, ▶Section 2.6.3. The complex forma-
tion is only possible if the chromophore or its substituents can form coordinative bonds
with the metal. Suitable groups are hydroxy and oxo groups, carboxylic acids, and diazo
groups; common structural elements are [19, p. 260]:

These elements are realized mainly in anthraquinones and azo compounds. Examples
are the tridentate azo dyes:
520 � 5 Dyes for writing, painting, and drawing

Based on the coordination numbers 3, 4, and 6 of the metals and the bidentate and tri-
dentate dye ligands, we can expect dye-metal complexes in the ratio of 1:1 and 2:1 [706].
To fully achieve the coordination number of the metal, secondary ligands such as water,
hydroxyl anions, halide anions, or acid groups are applied.

Adhesion
The dye-metal complex can be fixed via bonds between suitable groups of the substrate
and the metal, which, depending on the metal and the binding partner, have a more
covalent, coordinative, or ionic character, as illustrated by some examples of 1:1 and 2:1
complexes (dye to metal ratio):

Meanwhile, there is no confirmation of this plausible theory, and in the case of some
2:1 complexes, the necessary free valence at the metal would not be present. It is, there-
fore, assumed that the adhesion is due to ionic bonding, dipole-dipole interactions, and
dispersion forces already effective in the underlying free dye [19]. In addition, the com-
plex has a significantly reduced diffusion rate compared to the free dye since its size
means that more interactions with the substrate occur than per isolated dye molecule.
Furthermore, the complex is physically trapped due to its sheer size.
For staining, two procedures have been established. They differ in the order the
reagents are added:
– Traditional staining process: soaking of the substrate with the mordant, then adding
the dye
– Staining process with mordants: dyeing of the substrate with suitable dyes, then
adding the mordant

Commercially, the second method almost exclusively is used today. The applied method
achieves dark colors such as black.

Traditional staining process


For staining, according to the first method, the dyeing material is immersed in a solution
of metal salt such as alum, AlCl3 , or FeCl3 , and saturated with it. In this process, the
5.5 Mordant dyes (metal complex dyes) � 521

metal cations form salts or complexes of more or less ionic bonding character with acidic
groups (hydroxyl, carboxylic acid, or sulfonic acid groups) of the fibers [207, p. 484], [208,
p. 518], [209, Chapter 34.6.5]. Subsequently, the mordant is fixed with hot water vapor,
and any valencies of the metal cations that remain free are hydroxylated. In addition
to the metal hydroxides that are ionically or complexly bound to the fiber, metal oxide
hydrates also form, which are distributed in the fiber. Then the dye solution is added
to the stained dyeing materials. Dye anions replace hydroxide ligands in equilibrium
reactions, and subsequently, the insoluble metal complex is formed on and in the fiber.
Ionic and complex bonds anchor the complex to the fiber:

After-chroming
This process employs metal salts as mordants. In the case of chromium mordant, it is
called “after-chroming”, but the term got used for this process in general. In this pro-
cess [19, Chapter 13.7.3], [20, p. 85], the substrate is first dyed with a suitable dye (e. g.,
acid dyes for wool or disperse dyes for nylon). In the following step, by adding the mor-
dant, the metal complex is created:

In this order, it is unlikely that the metal cations will form many ionic bonds with the
substrate, and thus contribute to fixation.
As the name suggests, CrIII salts are very popular in practice as mordants because
intense dark shades (blue, brown, black) are obtainable. Dichromates are used as metal
salts; CrIII is derived from them by reducing groups in the substrate or sodium thiosul-
fate. Since contamination of wastewater and dyeing materials with the toxic hexavalent
522 � 5 Dyes for writing, painting, and drawing

CrVI could be possible, the share of chromium mordants slowly decreases in countries
with environmental protection regulations. In the course of replacing chromium and
cobalt with nontoxic metals, iron complexes with red, purple, blue, and black colors
have been found [707, 708].

5.6 Cationic dyes


Cationic or basic dyes are water-soluble compounds that are electrostatically fixed to
anionic groups in the substrate [13, 19, 20], [203, keyword “cationic dyes”]. Their colors
are often brilliant and intensive.
The first cationic azo dye, vesuvin, was described as early as 1863 and was obtained
by coupling m-phenylenediamine to m-phenylenediamine. Chrysoidin followed in 1875.
Mauvine is significant for the development of the dyeing industry, and thus for today’s
large chemical corporations since its discovery opened up the era of synthetic dyes. The
emergence of synthetic fibers also increased the value of cationic dyes since their main
application is the dyeing of nylon, anionic polyacrylics, and polyester fibers or leather.
They show a high affinity for wood-containing paper, so they are frequently used for pa-
per, inks, and stamping inks. An advantage is their color brilliance, but their lightfastness
is usually low, making them only suitable for cheap wood-containing papers, wrapping
or packaging materials, or for applications where they are not exposed to light for long
periods.

Chemistry
The variable chemistry of cationic dyes applies hemicyanines from diazo compounds,
triarylmethines, cyanines, thiazines, oxazines, and acridines. The chromophores carry
cationic centers, e. g., ammonium groups, and form salts with anions of simple acids
(halide, acetate, oxalate, or sulfate). Due to a large number of coloring structures, ba-
sic dyes cover the entire color spectrum. They are named “basic dyes” because the dye
cations form insoluble “basic” precipitates dye⊕ ⊖ OH with alkali hydroxides.

Adhesion
Cationic dyes form ionic bonds with anionic positions of the substrate, e. g., carboxyl
groups. Their suitability for paper dyeing (▶Section 5.2) is due to their high affinity
for wood-containing and unbleached wood-free aqueous paper pulps. Lignin and acid
hemicelluloses in these pulps exhibit a high anion activity that binds the cationic dyes
during production. However, their affinity to bleached paper is low, and auxiliaries are
needed, which are admixed to the paper pulp as fixatives (anionic polymers, ▶p. 602) or
added as an anionic component during sizing, ▶p. 605.
5.6 Cationic dyes � 523

Examples
Cationic dyes in inks for felt-tip pens are often based on the triphenylmethane skeleton,
▶Section 4.8.1 at p. 370, less frequently on the diphenylmethane skeleton:

Azo dyes, acridines, and cyanines are also suitable ink dyes for felt-tip pens:
524 � 5 Dyes for writing, painting, and drawing

For coloring paper in pulps, BR1, BO2, and BBr1 are applicable; BR1 is also a suitable ink
for felt-tip pens:

5.7 Anionic or acid dyes


The group of acid dyes comprises water-soluble compounds carrying anionic sub-
stituents. These increase water solubility and impart ionic bonds to the substrate [13,
19, 20].
Appropriate substrates for acid dyes are fibers with amino groups (polyamides,
wool, silk) or other substituents that form cations by protonation, leather, and treated
paper. Due to their affinity for paper, they are used to a lesser extent for paper coloring
but more commonly in writing and inkjet inks. Acid dyes with low molecular weight are
unsuitable for cellulose; with a high molecular weight, they change into direct dyes.

Chemistry
The history of acid dyes began around 1876 with naphthol orange and later Eosin (AR87),
which is still a dye of red ink today. Acid dyes are based on all of today’s chromophores.
Azo chromophores cover the color range from yellow to red, purple, and brown; disazo
compounds achieve blue and black. The resulting large structures are already close to
direct dyes. Anthraquinone chromophores cover the range from purple-blue to green,
5.7 Anionic or acid dyes � 525

and copper phthalocyanine chromophores that of light turquoise. For lower lightfast-
ness requirements, triphenylmethines are used for blue to green and xanthenes for red
to purple.
Chemically, acid dyes are compounds with molecular weights of 300–1000 g/mol car-
rying anionic substituents (sulfonic acids, carboxylic acids). In particular, the sulfonic
acid group is readily introduced by sulfonation of the structure. The anionic groups in-
crease the solubility and form ionic bonds with the substrate, which in the case of acid
dyes of low molecular weight, achieve adhesion to the substrate.

Adhesion
The name “acid dye” is derived from the acidic environment of the dye bath, with pH
values of less than 2 (sulfuric acid) to more than 6.5 (ammonium acetate). The processes
involved in dyeing a protein fiber such as wool or a synthetic fiber such as nylon are as
follows:
– The amino acids of the proteins are present as zwitterions, according to their iso-
electronic points.
– Amino acids or the C and N termini of the nylon chains are converted in the acidic
dye bath into nondissociated carboxylic acids and ammonium salts of the added
acid.
– The acid anions are replaced by dye anions, which ionically adhere to the fiber.

Since cellulose has no cationic charges, acid dyes show no affinity toward untreated pa-
per made from plant fibers, ▶Section 5.2. Therefore, adhesion to paper depends on its
properties, which can be adjusted by a suitable paper pulp or coating composition. Pos-
sible fixatives (aluminum sulfate, cationic polymers) are presented on ▶p. 602, suitable
sizing agents (resin sizing) in ▶Section 6.7.4.
The ionic bonding described primarily contributes to the fixation of small dye
molecules. With increasing molecular weight and more extensive nonpolar regions,
acid dyes adhere to the substrate additionally via dispersion forces and dipole-dipole
interactions and transition seamlessly into direct dyes. ▶Table 5.5 shows the influence
of molecular weight.

Examples
The excellent adhesion of acid dyes allows them to become alternatives in the dyeing
industry for cases where the migration fastness of disperse dyes is too low in some
synthetic fibers. AY17/AR37/AB45 and AY29/AR57/AB40 are primary color triplets, with
526 � 5 Dyes for writing, painting, and drawing

Table 5.5: Influence of molecular weight Mr of acid dyes on various dye properties [19, Chapters 13, 13.4],
[20, p. 81].

Mr low Mr high

Substantivity for wool, nylon Low High


Diffusion in fiber High Low
▶migration fastness Low High
Hydrophobicity Low High
Washing fastness Low High
Adhesion Ionic bonding Hydrogen bonding, dipole-dipole
interaction, dispersion forces
Solubility High Low
Sulfonic acid groups 1–3 1

which knitwear and nylon tights can be dyed in all the required shades without aller-
gens. Some of them, such as AY17 and AR37, are used for inkjet and writing inks:
5.7 Anionic or acid dyes � 527

Another primary color triplet AY23/AR52/AB9 is used together with Food Black 2 in inkjet
inks:

The following dyes are examples of yellow, orange, red, and blue inkjet and writing inks.
Further examples (AY73 and AR87) were given with the triphenylmethane dyes, ▶Sec-
tion 4.8.1. AG27 and AV41 provide examples of green and purple acid dyes, respectively:
528 � 5 Dyes for writing, painting, and drawing
6 Structure of paint systems
The following chapters are dedicated to the actual artistic techniques. Colors are an es-
sential part of each technique, as varied as the techniques and their results. By “colors”
we do not mean the visible hues, but the paint systems, i. e., the paints (the color paste
essential for oil and acrylic painting), pencil leads, printing inks, or colored solutions to
create a work of art.
Colors differ depending on the type and technique of painting, but there are never-
theless similarities in their structure. We will therefore address general characteristics
here, and in subsequent chapters, we will focus on the specifics of certain painting tech-
niques. More can be found in [47, 74, 75, 97, 98, 117–122, 187, 188], Tadros [221] explicitly
considers the role of colloids in paints, and Tadros [222] describes the theoretical basis
of wetting processes.
Paints are generally composed of several components:
– colorants (pigments, sometimes dyes) and fillers
– binders (film former)
– solvents
– auxiliary substances

The latter support the main constituents in their function or perform particular tasks.
They are only contained in paint to a minimal extent in terms of quantity. Auxiliary
substances can be:
– wetting agents and dispersants
– film-forming aids
– thickeners and rheology modifiers
– defoamers, siccatives, preservatives

The most critical components, also in terms of quantity, are colorants, fillers, binders,
and solvents. Pigments that give the paint color, body, and materiality can vary depend-
ing on the technique, but we can usually employ a pigment in several techniques. We
have discussed colorants in ▶Chapters 3, 4, and 5. ▶Section 1.6.8 already made us aware
that pigments and fillers can only have the desired opacity if their refractive index is
sufficiently high concerning air, water, and binder.
In addition to paints, media and binders are essential products of artists’ supplies.
Partly, they share some components with paints, but in a different composition since
they provide only one aspect of a paint:
– binder
– solvent
– defoamer, preservative, film-forming aid, matting agent, gloss enhancer

[Link]
530 � 6 Structure of paint systems

6.1 Binders

Binders permanently fix pigments to the support or ground by forming a film that traps
them (and also most of the other paint components). The film can be formed physically
by solvent evaporation and retention of the film-former on the substrate or chemically
by cross-linking the molecules of the film former. Binders, however, also provide essen-
tial optical properties of a paint system along with the film. These include film thickness
(▶ deep light, transparency, chalky-matte appearance), gloss, and color depth. In addi-
tion, the binder determines usage properties such as film hardness, film resistance, and
abrasion resistance.
However, binders can significantly influence the processing properties during the
application, depending on their proportion in the paint. These properties include, above
all, the open time or drying time, i. e., the time during which the paint can still be ma-
nipulated or processed. Whereas paints for coatings and paints should only be “open”
during the application itself, and then rapid drying is desired, the artist requires a wide
range of fast, medium, and long-drying paints. Following this essential requirement, a
wide range of binders exists for artists’ materials, ▶Figure 6.1. Another critical factor is
the viscosity, which is a binder’s property. If necessary, it must be modified by adding
solvents (thinners) or thickeners.
After a historical overview and an outlook on the problems of binder analysis, we
will deal with the physical-chemical fundamentals of drying and the forces that lead to
bonding.
Binders are macromolecular by nature to provide the desired fixing and protec-
tive film around the pigments. Usually, they have molecular weights in the range of
500–30000. Binders with low molecular weights have the advantage of lower viscosity
during processing but must polymerize on drying to form a durable film. ▶Figure 6.1
roughly indicates the mass range in which the essential binders can be classified: Most
binders are polymeric by nature; drying oils and oil-modified alkyd resins are low
molecular weight but polymerize on drying. Resins are low molecular weight com-
pounds that do not polymerize on drying.

6.1.1 History

Binders have a rich past dating back to Neolithic times [111, 117, 120, 128, 921]. Some
milestones of chemistry are indicated in the following outline:
6.1 Binders � 531

Figure 6.1: Overview of the leading chemical binder classes. Blue background: chemically drying. Light
gray: predominantly physically drying. White: purely physically drying [100]. Left: low molecular weight
compounds. Right: high molecular weight compounds.
532 � 6 Structure of paint systems

47000 BC Possible casein paint (South Africa, ocher dissolved in milk)


20000 BC Natural fresco (calcium bicarbonate-bound ocher drawings), egg white, blood
7000 BC Fresco secco (Çatalhöyük)
3000 BC Gums, such as gum Arabic and cherry gum (watercolors, fresco secco, soot ink (carbon))
3000 BC Proteins (glues, such as skin and fish glue), soot ink (carbon), papyrus
2000 BC Fresco buono (Crete)
1500 BC Proteins (glue, egg) (watercolors and tempera, fresco buono and secco), (vegetable gum with
honey, casein, starch, tragacanth, glue, egg white, egg tempera), glazes (to absorb colorants)
1000 BC Proteins (egg tempera)
500 BC Resins, fresco, encaustic (panel painting on wood)
200 BC Parchment
0 Nut oil, gums, milk
100 Paper
600 Earliest mention of oils with pigment
800 Book illumination, gum and albumen bound
850 Egg tempera (icons)
1000 Linseed oil, gum Arabic
1100 Early example of an oil binder in Denmark
1250 Egg tempera (duecento panel painting)
1400 Oil painting with turpentine oil in Early Netherlandish painting (distillation from Arabia),
walnut oil, linseed oil, watercolor (gum-bound), gouache (gum and glue-bound)
1500 Shellac (by India trade)
1500 Oil painting in late Renaissance art
1600 Lacquers
1700 Low-fusing glaze for onglaze paints
1877 Polymethacrylic acid (Sittig, Paul)
1901 Acrylic polymers (Röhm)
1905 Cellulose ethers (Suida)
1912 Polyvinylchloride and polyvinylacetate (Klatte)
1926 Polyvinylalcohol (Herrmann)
1930 Oil-modified alkyd resins (Kienle), acrylic dispersions

▶Figure 6.2 shows crucial phases from an artistic point of view. Early cave paintings
have been faithfully preserved in a fresco-like technique, though probably uninten-
tionally, by calcareous water. The pigment, which at first adhered to the rock wall only
by pure adsorption or inclusion in fat, was washed over by the calcium hydrogen car-
bonate contained in the dripping water and was therefore enclosed in a lime casing by
calcination:
6.1 Binders � 533

Figure 6.2: Rough chronological overview of the essential paint systems [72, 73, 117, 120, 128]. Top, blue:
paint systems. Bottom: binders.

The reaction proceeds similar to the setting of a fresco binder (▶Section 7.3), which con-
sists of slaked lime and also sets to lime:

This true fresco is deliberately used from the second pre-Christian millennium onward.
It is later perfected in the Hellenic and Roman periods before experiencing further flour-
ishing in the Middle Ages and Baroque. Initially, however, in the earliest human settle-
ments, loam/clay and gypsum plasters were used as wall coverings and painted with
watercolors in a fresco secco technique. Their binders are based on polysaccharides
(gums) from plant or tree saps (fig milk, gum Arabic) or glue from proteinaceous sub-
stances (egg whites, casein, skin, bone). The gum-bound paints are refined in modern
times to today’s watercolors and gouaches; the protein-bound paints are precursors of
today’s poster paints and other glue paints. Tempera was already developed in antiquity,
using aqueous (gum, glue) and nonaqueous (resin, oil) components as emulsion binders.
Gum, glue, egg, and lime remained the main binding agents until the Middle Ages.
Then, besides mural paintings, book illumination with watercolors emerged as a vital
form of artistic expression. The origins of European panel painting lie between the turn
of the first millennium and the Renaissance. Tempera with egg (egg tempera) is its dom-
534 � 6 Structure of paint systems

inant binding agent, as we can see from the typical appearance of numerous master-
pieces, e. g., of the early Italian painting of the Duecento and Trecento.
The addition of drying oils to egg tempera leads to oil painting. Within a few hun-
dred years, it became the dominating technique thanks to the Early Netherlandish paint-
ing in the fifteenth century and significantly changed the art’s appearance. Paintings
before 1505 were predominantly executed in egg tempera, and only isolated areas were
worked in walnut oil [929, 930, 932, 933, 935–942]. After 1500, artists used oil almost exclu-
sively in paintings. At the beginning, walnut oil was employed, later rather linseed oil,
until around 1700, linseed oil (occasionally poppyseed oil) was predominantly applied.
Higgitt and White [931] illustrate the differentiated parallel use of both media in Italy,
e. g., oil for glazes. Detailed studies of individual artists such as Dürer and Tintoretto
are given in [922–924, 943]. The latter shows that, e. g., Van Dyck used walnut oil and
linseed oil side by side because of their respective advantages; the lighter and less yel-
lowish walnut oil for sensitive color tones such as white, yellow, or blue; for colors such
as black, which dried with difficulty, he also used stand oils or boiled oils.
A unique feature of the eighteenth and nineteenth century was megilp, painting
agents consisting of a mixture of drying oils, lead compounds, and mastic varnish. They
showed a pronounced thixotropic behavior [934]. While they could be painted smoothly
for glazes, they were also suitable for impasto, which requires a paint that is buttery and
“stands up.” Since they led to severe cracking and yellowing, they were applied more and
more rarely. Well-known users were Reynolds, Turner, and Wilson.
Oil painting retained its supremacy until the last century. In the field of coating
technology, from the nineteenth century onward, the oils were ester-modified and led
to alkyd colors, which then in the twentieth century became popular also in the panel
painting. Alkyd paints, however, could not displace oil binders.
So far, the last development impulse created acrylate-based dispersions, which were
initially also used for coatings. From around 1950, acrylic dispersions became available
for artistic painting. They rapidly advanced to dominant binders besides oils since they
were inexpensive, water-soluble, low-odor, easy to process, and adherent to many sur-
faces. However, since acrylic film differs considerably from oil film, it remains to be
seen to what extent the painting properties and the considerably different appearance
of acrylic paintings will be able to compete with oil painting (including its advantages
and disadvantages) in the long run.

6.1.2 Binder analytics

When analyzing art objects, we encounter particular problems in the case of binders,
which often make it difficult or impossible to determine the exact agents used. Most col-
orants have a characteristic structure that facilitates analysis or at least classification.
Problematic colorants are usually bio-organic, such as lakes, which fade, degrade, and
6.1 Binders � 535

have similar structures, so individual representatives are difficult to identify. These dif-
ficulties are even more pronounced concerning binders before the last century: they are
primarily of natural origin, have complicated or polymeric structures (proteins, polysac-
charides), belong to a class of substances with numerous similar representatives, and
are often modified or degraded over long periods. A detailed analysis has only recently
become possible using methods such as chromatography (GC, HPLC, TLC) and mass spec-
troscopy. Furthermore, one-dimensional and two-dimensional NMR spectroscopy has al-
lowed for significant advances; here, 1 H-NMR and 13 C-NMR are especially suitable meth-
ods.
[101, 102, 921, 925–928] give insights into the analytics of binders and show possible
solutions. Since individual substances are hardly representative, characteristic ratios of
dominant substances or classes must be determined for each binder. Samples have then
to be compared with these ratios.
For oil-based binders, information can be obtained on the relative amounts of fatty
acids in the oil’s glycerides. Bonaduce et al. [928] list examples often used to identify
a drying oil. These are the ratios palmitic acid/stearic acid, azelaic acid/suberic acid, or
azelaic acid/sebacic acid. However, the figures obtained strongly depend on the pigments
employed and the painting’s age, i. e., on processes during the oil’s drying. Suitable for
identification are also hydrolysis or oxidation products. In the case of egg tempera, in-
stead of the fatty acid ratios, the amount of azelaic acid can be determined compared to
palmitic and stearic acid. Egg-containing binders also have cholesterol content.
Nondrying lipids indicate binders such as animal glue. Glues, as well as tempera,
are characterized by their protein content. The relative amounts of amino acids then
permit conclusions about the exact nature: high contents of hydroxyproline or glycine
suggest collagens, the absence of hydroxyproline in the case of high proline and leucine
contents indicates casein, the presence of leucine and aspartic acid point to egg or egg
white. Unfortunately, glues are binders easily destructible.
Identifying plant gums such as gum Arabic, gum tragacanth, or gums from pome
fruit in painting specimens is challenging and doubtful. Therefore, only a few investi-
gations on art objects are available so far. Plant gums, e. g., the three gums mentioned
before, are effluxes from the endosperm of some seeds, which consist of high molecu-
lar weight polysaccharides. The main components of polysaccharides are aldopentoses,
aldohexoses, and uronic acids. The presence or absence of key sugars indicates the ori-
gin of gum. After the hydrolytic breakdown of polysaccharides into monosaccharides,
GC/MS can be employed to analyze the gum base. However, the analytical results of-
ten show deviations from reference materials. Little is known about aging and related
changes in polysaccharides, including their interactions with the carrier materials, envi-
ronmental influences, or bound pigments. In [927], patterns of key sugars are proposed
that are stable to aging and pigment effects, thereby providing a decision tree based on
specific sugars such as mannose, xylose, or fucose; in this way, the underlying gum can
be identified in proper order.
536 � 6 Structure of paint systems

Acrylic binders [99] can be clearly distinguished from binders of natural origin be-
cause they contain many synthetic components. These include, e. g., polyethylene gly-
cols, and residual monomers of polymerization.

6.1.3 Drying of binders

To understand binders, we need to imagine a painting’s drying process. Depending on


the paint system, this involves processes that pursue two objectives:
– removal of the solvent that was necessary to transform the powdered pigments into
paintable colors
– creation of a solid bond between the pigments and the support or ground, e. g., by
embedding the pigments in a transparent binder film or by glueing

We can achieve both objectives in two ways by:


– chemical drying
– physical drying

In chemical drying, after or during the solvent’s evaporation, the binder polymerizes
and forms a chemically different solid substance binding the pigments to the substrate.
In most cases, a crystal-clear polymer film forms, embedding the pigments. An example
is oil paint, in which linseed oil cross-links to an insoluble polymer when exposed to air
and light.
In physical drying, the solvent evaporates. However, the pigments and a dissolved,
nonvolatile binder remain. The binder adheres to the substrate, binds the pigments to
the substrate, or encloses them in a film-like layer. The binder does not change chemi-
cally in the process. Resinous painting agents such as dammar varnish provide an exam-
ple. The solvent (turpentine oil) evaporates rapidly, and the nonvolatile resins remain on
the support or ground, binding the pigments. Some paints even dispense with binders
altogether, relying on the pure adhesive forces of pigments.
Both variants are often combined to achieve better application properties of the
system. Thus, e. g., in the technique of the Old Masters, on oil paint layers, a white high-
light with aqueous egg tempera is applied. This tempera layer dries physically, its wa-
ter evaporating quickly. As a result, this layer can be painted over, and its oily fraction
dries chemically together with the other oil layers over time. In addition to application-
related advantages, there are also advantages to the painting technique. For example,
the distinct oily and watery systems lead to sharply defined, hair-thin drawings, as we
can impressively see in fur and hair in Dürer’s paintings.

Chemical drying
Chemically drying binders form the basis for some of the best-known traditional paint-
ing methods. These include:
6.1 Binders � 537

– drying fatty oils


– calcium hydroxide
– oil-modified alkyd paints

Drying fatty oils comprise, e. g., linseed oil, walnut oil, or poppy seed oil. In this classical
oil paint system, the oil molecules cross-link covalently under oxygen absorption. As a
result, they build a three-dimensional network in the form of a thick glass-clear film,
▶Section 7.4.
Calcium hydroxide is the binder of the genuine fresco. The chemical drying process
is carbonatization, the formation of calcium carbonate with carbon dioxide from the air,
▶Section 7.3.
Oil-modified alkyd paints are a new development. They dry chemically due to the
oil content.

Physical drying
Physically drying binders have also been in service since historical times:
– plant gums such as gum Arabic or gum tragacanth as a base for antique tusches and
inks, water-based colors, and modern watercolors
– animal glues (proteins), e. g., glutin glues (collagen) from skins and bones, egg white
from egg, casein from milk for inks and watercolors
– resins such as dammar or mastic as a component of varnishes and resin oil paints
– acrylic dispersions as the latest extension of painting agents for acrylic paints

Over time, binders of animal glues polymerize through chemical changes, cross-link,
and thus, in some cases, undergo chemical drying.
In contrast to chemical drying, the binder does not change in purely physical dry-
ing. The processes that lead to a stable film or an adhesive layer are based on secondary,
intermolecular interactions such as electrostatic and van der Waals forces, which we
consider in detail in the next section. Most binders in this category are polymers, which
provide a large number of such weak, secondary bonds; for this reason, extended, insol-
uble polymer networks form. In ▶Figure 6.3, we see essential steps of physical filming:
– In the dry binder, the molecules form polymer coils, ▶Figure 6.3(a).
– In an ideal solvent, the polymer coils disentangle and the molecules are stretched
until they are completely enveloped by the solvent. In contrast, in an unsuitable
solvent, the polymer coils predominantly remain intact, ▶Figure 6.3(b).
– Evaporation of the solvent starts the film formation:
– The formerly isolated binder molecules mutually approach, either by their ther-
mal migratory motion, or by the solvent’s evaporation, increasing the binder
concentration, ▶Figure 6.3(c). In case of a good solution, large regions of the
stretched molecules can optimally overlap, fixing themselves via intermolecu-
538 � 6 Structure of paint systems

Figure 6.3: Schematic process of physical drying of polymeric binders. White box: support. (a): The pow-
dered binder comprises isolated polymer coils. (b): In solution (blue), polymer coils unfold and binder
molecules stretches. (c): The expanded molecules interact due to evaporation and thermal movement.
Evaporation increases the binder’s concentration and, therefore, interaction probability. (d): Partially crys-
talline or intermolecularly linked regions form in the film.

lar interactions. For a poorly dissolved binder, the individual polymer coils only
slightly interact.
– Parts of different polymers become entangled and interlock tightly by inter-
molecular forces, forming quasi-crystalline regions, ▶Figure 6.3(d). The forma-
tion of hydrogen bonds provides additional stabilization. As we will see, such
a process is irreversible once certain energy thresholds have been overcome,
thus leading to films that can no longer be dissolved. The polymers can similarly
interact with the support so that the evolving cross-links also become fixed to
the support, forming a solid film.
For a poor solution, the individual polymer coils remain loosely coupled. A frag-
mented film of inferior stability results, which can readily disintegrate into in-
dividual coils, ▶Figure 6.3(d).

6.1.4 Bonding types in the world of colorants

We have already addressed the role of secondary bond types in physical drying, which
leads to forming a compact film. However, now is the time to explore the mechanisms in
the pigment/dye-binder-substrate system responsible for film formation and adhesion
of pigment or dye. We will mainly encounter noncovalent bonds, ▶Table 5.1. They do not
6.1 Binders � 539

link the atoms of a molecule, but molecules among each other and have low strengths.
Most interactions of colorants with their environment are therefore weak, ▶Table 6.1.

Table 6.1: Relative strengths and energies of secondary bond types in colorants and binders.

Bond type or strength Relative strength [19, pp. 42] Energy [kJ/mol] [220, p. 35]

Ionic bond 700 590–1050


Covalent bond 400 60–700
Hydrogen bonds 40 50
Dipole-dipole 5 20
Dispersion force 5 42
Dipole-induced 2

Ionic bonding
Ionic bonding depends on both bonding partners’ electrostatic attraction of opposing
electrical charges. Typical examples from chemistry are crystalline salts, whose crystal
lattice is built up of anions and cations. Interesting for our topic is ionic bonding when
used to bind basic, acid, and mordant dyes to the substrate, ▶Chapter 5.

Noncovalent or van der Waals interactions


The very weak noncovalent, attracting van der Waals interactions also act between un-
charged and nonpolar atoms and molecules. Since they diminish rapidly with increasing
distance between the partners, typically with the sixth power, we usually cannot notice
their effects on a large scale. They occur wherever particles cluster together and we ob-
serve adhesion effects, i. e., effects associated with spatial proximity of molecules (on a
small scale) or particles and extended surfaces (on a large scale).
– Fine powders tend to form clusters as the particles aggregate.
– Dyes and geckos adhere to their support with the help of their extended molecular
or body surfaces.
– Glues, which do not react chemically, adhere to the surfaces through noncovalent
interactions.
– Physical film formation in proteins or latex dispersions is achieved by aggregation
of the binder molecules.

Under the term “van der Waals interaction,” we summarize three very similar inter-
actions, although frequently no clear distinction is made between them (as it is often
unnecessary):
– the dispersion interaction (London force) between two polarizable molecules, i. e.,
two induced dipoles
– the Keesom or dipole-dipole interaction between two dipoles
540 � 6 Structure of paint systems

– the Debye interaction between a dipole and a polarizable molecule, i. e., an induced
dipole

Often the term van der Waals force refers to the London dispersion interaction, which
is also the dominant one of the three forces.

Origin of dipoles
Common to all three forces is a charge separation (dipole formation), which occurs in
both partners and induces dipole moments.
Permanent dipoles are evoked by the different electronegativities of the atoms, and
thus charge densities in the molecule, e. g., in compounds that contain oxygen or nitro-
gen (hydroxyl and amino groups, carbonyl groups, nitriles). A well-known example of
dipole-based interactions is the hydrogen bonds.
Temporary dipoles emerge because electrons (considered in the particle picture) are
localized in very tiny time intervals, and a charge separation occurs femtosecond by fem-
tosecond, even if the molecule is neutral on time average (spontaneous polarization).
They occur, e. g., in the interaction of nonpolar proteins in casein paints (▶Section 7.5.3)
or the adhesion of dyes to nylon fibers and polyethene films (▶Chapter 5).
Dipoles of both types also induce a charge separation and a dipole moment in their
neighborhood: the original dipole A generates an electric field by which particle B is
polarized. The resulting dipole B also generates an electric field, which interacts with
that of A and leads to mutual attraction if B aligns itself to A and the fields can overlap
constructively.

Orbital interpretation of the interaction


An interpretation of the van der Waals interaction in the framework of the orbital in-
teraction theory is outlined in [213, Chapter 3]: the interaction of a spatially localized
occupied orbital of one partner with an unoccupied, spatially extensive orbital of a sec-
ond partner leads to an energetic lowering of the occupied orbital, and thus to an energy
gain. With decreasing distance, the lowering progresses further until occupied orbitals
of both partners interact and develop a repulsive force.

Range of interaction
In the molecular range, van der Waals interactions can be described by a potential that
decreases with the sixth power of the distance between the particles, i. e., it decreases
rapidly and has a range of a few nanometers:

k
Va (r) = − (6.1)
r6

In the Lennard–Jones potential, these interactions represent the attracting term.


6.2 Solvents � 541

In the order of magnitude of colloidal particles (a few hundred nanometers), the


attracting forces should already have decreased to an imperceptibly small value. How-
ever, an important effect occurs here: every single atom can interact to a certain extent
additively with all atoms of the partner particle so that we observe a focusing of the
dispersion force that is called long-range van der Waals force or Hamaker force. Inter-
estingly, it only decreases approximately quadratically with distance [218, p. 268]:

k
VaH (r) = − (6.2)
r2
Its range extends to the order of 100 nm, so this force contributes decisively to the dy-
namics in dispersed systems; as we will see later in the example of acrylic dispersions,
▶Section 7.9.1.

Hydrophobic interactions
The term “hydrophobic interaction” is often used to describe an entropic phenomenon:
The tendency of hydrophobic groups to agglomerate in water or aqueous solutions and
the resulting reduction of contact area between the hydrophobic groups and the wa-
ter.
The cause of this aggregation can be described entropically. A thin layer of water
molecules evolves at the interface of a nonpolar particle to the water. They are more
highly ordered than the free water since all their hydrogen bonds are directed toward
the free water. This state implies a low entropy and a limitation of the mobility of these
water molecules. Nonpolar molecules that collide with each other through random ther-
mal motion reduce the interfacial layer, release some previously ordered bound water
molecules, and the entropy of the solution, including the aggregates, increases. The pro-
cess is irreversible since entropy can only ever increase in closed systems. As a result,
more and more particles aggregate, and the aggregates grow in size.
In addition, van der Waals interactions between the nonpolar particles occur dur-
ing aggregation, leading to an enthalpy gain. We note, however, that this effect becomes
strong only when the particles are very close to each other. The real driving force is
the entropy gain due to the release of solvent molecules from the hydrophobic inter-
faces.

6.2 Solvents
Solvents are only necessary during the application phase and escape after applying the
paint as it dries. Together with the binder, they provide specific processing properties.
Viscosity is the eminent property determined primarily by the binder but can be mod-
ified by adding solvents or thickeners. We will discuss solvents in individual paint sys-
tems if necessary.
542 � 6 Structure of paint systems

6.3 Wetting agents and dispersants, grinding paints


Wetting agents and dispersants play an essential role in producing or grinding paints
and their storage. They support the wetting of pigments with binders and stabilize the
pigment dispersions obtained by preventing flocculation, aggregation, or sedimentation
of pigments. As a result, we obtain an even surface of the painting layer with uniform
gloss, transparency, and other optical properties.
Anyone who, as a beginner, has ever tried to manually mix a color probably re-
ceived a hint concerning the importance of grinding. The author bore this wisdom in
his mind when trying to produce oil paint. Stirring the pigment into the linseed oil was
unproblematic, and soon the author had produced a thick paste that looked like oil paint,
smelled like it, and could well be picked up with a brush. To his horrified astonishment,
however, the color on the canvas literally turned to dust and left only a dirty stain on it.
What had happened? Stirring the pigment into oil had produced a seemingly thick
paste but had not created any bond between the pigment and the oil, i. e., no bond be-
tween the binder and the pigment particles had been created. Chemistry provides ex-
planations and tools to produce high-quality colors, but first, we need to take a closer
look at the elementary steps of grinding [190, Chapter 4], [221, 222], [185, Chapter 4.3.2]:
– step 1: wetting of the pigment powder
– step 2: fragmenting of aggregates and agglomerations and dispersing of the primary
particles
– step 3: long-term stabilizing of the dispersed particles

Initially, the pigment is present as a powder in the air. However, for the successful pro-
duction of color pastes, the powder must be thoroughly wetted with the binder, i. e., the
air-pigment interface must be replaced by a water-pigment one.
The adsorption energy of the solvents is often not sufficient to spontaneously pro-
vide the energy required to form the new interfaces. Therefore, we bring mechanical
energy into the system and support the wetting by grinding.
Then, even a pigment supplied by the manufacturer as a powder consists of large ag-
gregates of molecules, viewed microscopically. In the production of pigments, we do not
directly obtain isolated molecules since these tend to form compact aggregates of many
molecules. This tendency is particularly evident in planar organic molecules, which
stack flat on top of each other. The weak interactions holding the aggregates together
generate large forces because of the spatial proximity. Consequently, the aggregates are
very stable. However, weak interactions are also active at the supramolecular level. Pig-
ments that are not synthesized directly but are extracted from minerals and are finely
ground form larger complexes, the agglomerations, through weak interactions. These
are not two-dimensional but are linked via corners and edges of the particles and show
weaker cohesion. Instead, they have a large number of cavities containing air bubbles.
Their interfaces to air, which must be wetted, are thus considerably more extensive than
those visible at first glance.
6.3 Wetting agents and dispersants, grinding paints � 543

By grinding, we disintegrate these agglomerations and aggregations, which are al-


ready wetted by the solvent, into small fragments (primary particles). The mechanical
energy that we supply by grinding is largely retained as surface energy in the greatly
enlarged surface of the pigment dispersion.
Finally, dispersing is the process of countermanding the attractive forces within the
aggregations and forming durable new interfaces between the pigment and the solvent.
A dispersion tends to flocculate, since the dispersed pigment particles tend to reaggre-
gate. Consequently, we must repress this tendency and stabilize the pigment dispersion
to produce a durable paint paste.

6.3.1 Wetting agents

Additives can support all steps: Wetting agents such as surfactants, reduce the interfacial
tension, and thus the energy required to form the interface with the solvent in step 1.
In addition, dispersants and stabilizers help separate the primary particles obtained in
step 2 during grinding and keep them from reaggregating. In the long term, they thus
maintain the dispersed state in step 3.

Theoretical treatment
The theoretical treatment of the wetting process shows that the essential quantity nec-
essary for good wetting of the pigment is the contact angle θ. θ is the angle at which a
drop of the binder forms with the pigment surface (▶Figure 6.4), the equation of Young
applies [185, Chapter [Link]], [221, Chapter 3], [222]:

γp − γps
γp = γps + γs cos θ or cos θ = (6.3)
γs

The γp , γs , and γps are the surface tensions of pigment (p) and solvent (s) or the interfa-
cial tensions between both, respectively. The smaller θ is, the better the wetting: Spon-
taneous wetting occurs for θ = 0, and the solvent spreads spontaneously over the entire
surface of the pigment, ▶Figure 6.4(a). For 0∘ < θ < 90∘ , the pigment wets well, ▶Fig-
ure 6.4(b) left. For 90∘ ≤ θ < 180∘ , the pigment wets poorly, and the solvent beads off,
▶Figure 6.4(b) middle and right. In the case θ = 180∘ , the pigment is not wettable at all;
the solvent forms closed drops, ▶Figure 6.4(c).
Spontaneous wetting occurs because the system gains energy. We can divide the
work required to wet or disperse a pigment with the surface A into three steps and cal-
culate for each the energy conversion, ▶Figure 6.5. In the adhesion, the pigment surface
comes into contact with the solvent and binder mixture. The immersion describes the
complete dipping of the pigment into the liquid. The spreading describes the detachment
of the pigment from the interface with air. The energy expended for each of these steps
can be described in terms of the respective interfacial energies:
544 � 6 Structure of paint systems

Figure 6.4: Influence of the contact angle θ on the wetting of a pigment surface (white box) with a solvent
(blue).

Figure 6.5: Model describing three steps in the wetting of a pigment (white box) by a solvent
(blue) [190, Chapter 4]. (a): Adhesion (contact between the pigment and solvent). (b): Immersion (com-
plete dipping of the pigment in the liquid). (c): Spreading (detachment of the pigment from the interface
with air).

Wa = γps − (γs + γp ) = −γs (cos θ + 1) (6.4)


Wi = 4γps − 4γp = −4γs cos θ (6.5)
Ws = (γps + γs ) − γp = −γs (cos θ − 1) (6.6)
Wd = Wa + Wi + Ws = −6γs cos θ (6.7)

We see that adhesion always occurs automatically (Wa < 0, energy gain), immersion
only when θ < 90°, and spreading only when θ = 0°. In the other cases, much effort is
necessary to achieve complete wetting. However, we also recognize that this situation
can be improved by reducing θ. Using ▶equation (6.3), we see that γs must be reduced
for this purpose:

⏟⏟⏟γ⏟⏟p⏟⏟ = ⏟⏟γ⏟⏟⏟
ps⏟⏟ + ⏟⏟⏟γ⏟⏟s⏟⏟ cos
⏟ ⏟⏟⏟θ⏟
⏟⏟ (6.8)
=const ≈const changeable should: change from −1 to +1

In the equation, γp is a constant that depends on the pigment, and γps is approximately
constant if we assume that a wetting agent essentially changes the properties of the
solvent. Both parameters cannot be changed or can only be changed at the expense of a
modification of the chemical structure of the pigment. To change θ from 180° to 0°, i. e.,
6.3 Wetting agents and dispersants, grinding paints � 545

to change the cosine from −1 to +1, we have to decrease γs , the surface tension of the
solvent. Using ▶equation (6.7), we can also say that we obtain a smaller contact angle
for the same dispersion work by this measure.
Unfortunately, we cannot independently change θ or γs as purely energetic consid-
erations suggested. They are also related to the depth l, to which the solvent will pene-
trate in a given time. l represents the solvent’s penetration into the cavities of the pig-
ment powder. The equation of Washburn relates l to high pore size r, low viscosity η,
and high surface tension γs :

k ⋅ r ⋅ γL cos θ
l=√ (6.9)

As a result, although we could now take a coarse pigment powder, we would have to find
an acceptable compromise between good wetting and fast wetting, as is often the case
when formulating paints and inks.
Reducing the surface tension of a solvent is the typical task of a surfactant. Due to the
similarity in structure, dispersion stabilizers can also act as wetting agents, significantly
the sterically active dispersants. Frequently used are, e. g., Aerosol OT® or branched
sodium dodecylbenzenesulfonate.

Surfactants
We divide surfactants or emulsifiers according to their charge into anionic, cationic, or
nonionic compounds [203, keyword “surfactants”], [966]. Cationic surfactants do not oc-
cur in the painting environment. However, we will see examples in the chapter on the
composition of inks, ▶Section 8.5.1 at p. 813. Anionic and nonionic compounds, on the
other hand, are frequently used in formulations for paints [185, Chapter [Link]]. Gen-
erally, they exhibit a chain-like structure with one hydrophilic and a hydrophobic end,
each of which shows a high affinity for either water or the polymer or pigment particle:
546 � 6 Structure of paint systems

At suitable concentrations, surfactants completely envelop the particles in a hedgehog-


like form (▶Figures 6.8(a) and 6.9(a)), and thus support the wetting of the particles. At the
same time, they can help keep the particles in the dispersed state: ionic surfactants, due
to their electrical charge, and nonionic surfactants, due to their steric claims, prevent
the particles from forming larger units. In the next section, we will address this effect to
function as a dispersant.
As ionic surfactants, sulfates and sulfonates of polyethylene glycols etherified with
fatty acid alcohols or alkylaryl sulfonates are applicable. Sulfonate groups are also intro-
duced with the help of succinic acid esters (sulfosuccinates). Linear or branched higher
alcohols (C7−18 ) are employed. Counterions are alkali, ammonium, or alkanolamines (tri-
ethanolamine):

For nonionic surfactants, similar substances are employed. Instead of the sulfate or sul-
fonate group, however, they carry hydroxyl groups:

The formerly essential alkylphenol ethoxylates are no longer used for environmental
reasons. Furthermore, betaines also belong to nonionic surfactants. These inner salts
are derived from the aminoacetic acid or the amino acid glycine:
6.3 Wetting agents and dispersants, grinding paints � 547

Common surfactants available commercially are listed as examples in ▶Figure 6.6. The
exemplary betaines are derived from stearic acid and “cocoic acid,” a mixture of higher
fatty acids derived from palm oil.

Figure 6.6: Examples of ionic and nonionic surfactants: alkyl and alkyl aryl sulfates and sulfonates, sulfos-
uccinates, and polyglycols. The last two compounds are nonionic betaines.

Surfactant production
Glycol alcohols (polyethylene glycols, PEG)
The solvents and surfactants mentioned here and the humectants mentioned later are
often established on glycol alcohols and their alkyl ethers. These compounds have fa-
vorable properties and can be easily prepared on a large scale, while their properties
can be varied in many ways. The starting point for the preparation is oxirane (ethylene
oxide), obtained from ethene in a catalytic process:
548 � 6 Structure of paint systems

The resulting ethylene oxide is used mainly in the production of polyester plastics such
as polyethene terephthalate (PET) and sizably as a solvent and antifreeze. Glycol alco-
hols are synthesized as follows:

If we apply alcohols instead of water (R is an alkanol or alkylphenol), we obtain ethylene


glycol alkyl ether or ethylene glycol alkyl phenyl ether. In both cases, the reaction can
usually be conducted so that only or predominantly monoglycols, diglycols, or higher
glycols form. In addition, if methyloxirane or propyl epoxide are applied instead of oxi-
rane, we also obtain significant polypropylene glycols.

Sulfates, sulfonates
Anionic surfactants based on sulfates and sulfonates are also cheap large-scale indus-
trial products. Alkane sulfonates are produced via sulfochlorination:

To prepare the alkylbenzene sulfonates, the corresponding alkylbenzenes are produced


by Friedel–Crafts alkylation from alkenes and reacted with fuming sulfuric acid:
6.3 Wetting agents and dispersants, grinding paints � 549

Alkoxyalkyl sulfonates can be prepared from alcohols and alkylglycols in an ether for-
mation reaction with chloroethanesulfonic acid:

Alkyl sulfates can also be readily produced from fatty alcohols by reaction with fuming
sulfuric acid, chlorosulfonic acid, or sulfur trioxide:
550 � 6 Structure of paint systems

6.3.2 Dispersants

Dispersants stabilize dispersions. Dispersions represent the important case of physically


drying binders that are not solutions. In general, dispersions are systems that comprise
finely dispersed particles in a homogeneous medium without forming genuine solu-
tions. Many well-known multiphase systems fall under the term: mist (liquid finely dis-
persed in gas), fume (solid in gas), suspension (solid in liquid), foam (gas in liquid), emul-
sion (liquid in liquid). In connection with artists’ paints, we encounter dispersions in
acrylic paints (polyacrylate suspensions in an aqueous solvent, ▶Section 7.9) and inks,
▶Chapter 8 and ▶Section 8.5.4 at p. 823. Hunter and Tadros [218, 221, 222] provide de-
tailed information about the exciting physics of disperse systems.
To understand how stabilization of dispersion paints and film formation work, we
need to know which processes take place in the dispersion when particles approach. As
a practical example, we will look at an acrylic paint dispersion in which the particles
are acrylate polymers. The noncovalent interactions involved were already addressed
in ▶Section 6.1.4. We assume that the particles are electrically charged, either already
given by the chemistry of the particles or by the addition of charged dispersing agents,
▶Section 6.3.3.
The interactions between the particles of the acrylic dispersion are described by a
theory called DLVO theory after the first authors Deryaguin, Landau, Verwey, and Over-
beek. A comprehensive summary is given in [218, Chapter 9], [219, 221]. ▶Figure 6.7(b)
shows the interaction potential Vt (r) between two particles of an acrylate-latex disper-
sion for three electrostatic repulsion forces of different strengths. Vt (r) is a function of
the distance between the particles. According to the DLVO theory, we can separate the
potential curve into several parts, which are easier to describe, ▶Figure 6.7(a):
– The dispersion or van der Waals forces have an attractive (negative) potential,
which in the form of the Hamaker force reaches a range of about 100 nm and shows
qualitatively an approximately inverse square dependence from the distance:
c
Va (r) = − (6.10)
r2
– The similar electric charge of the particles evokes a repulsive electrostatic potential,
which drops exponentially for spherical particles:

c′ −kr
Vr (r) = e (6.11)
k
6.3 Wetting agents and dispersants, grinding paints � 551

Figure 6.7: Total potential between two particles of an exemplary dispersion as a function of distance r.
Depending on the magnitude of the (electrostatic) repulsion, the primary minimum, the primary maximum,
and the secondary minimum can be distinguished. If the system is in the primary minimum, a nonresolv-
able film is present, while the secondary minimum describes the stable dispersion. The primary maximum
represents the potential barrier preventing coalescence and stabilizing the system. For this barrier to be
strong enough to prevent coalescence under surrounding conditions, the repulsion must be sufficiently
strong (light blue graph). If it is too weak, the system will coalesce on its own as it can constantly lose en-
ergy with smaller r (red graph).

– At very small distances, particles feel the Born repulsion VB (r) caused by the inter-
action of the orbitals with each other, which shows a very strong dependence on
distance and acts only on short distances:

c′′
VB (r) = (6.12)
r6
552 � 6 Structure of paint systems

Since quadratic functions increase more than exponential ones, the attractive potential
Va (r) is dominant for small and large r, i. e., the attraction between the particles predom-
inates in these ranges. If the repulsive potential Vr (r) is small, then Va dominates even
for any distance, resulting in a permanent attraction. The dispersion is, in this case, un-
stable and coagulates because it can gain energy (Vt becoming steadily more negative)
with an approach of the particles (r becoming smaller). The coagulation stops only with
the closest approach of the particles due to the Born repulsion becoming effective.
Only for a sufficiently strong, repulsive potential Vr , there is a specific range of r, in
which the repulsion predominates and builds up a barrier Em against a closer approach
of the particles, the so-called primary maximum. This barrier can be overcome by intro-
ducing energy, e. g., thermal energy from Brownian motion. Vt becomes negative again
at more considerable distances and forms the secondary minimum; at smaller distances,
the primary minimum.
What does the potential curve Vt (r) mean for acrylic dispersions? If particles ap-
proach each other sufficiently (r large), the van der Waal forces initially prevail until an
equilibrium is reached at (3) between the electrostatic repulsion and the attraction. This
point characterizes a stable dispersion.
At a further approach, the electrostatic repulsion prevails, and the energy increases
up to a maximum Em (2). This quantity represents the barrier against coalescence and is
a significant number that depends on the surrounding conditions. For example, at high
salt concentration, it is lower; at uncharged polymer particles, it does not exist.
Suppose we overcome this barrier by external forces and bring the particles
closer together. In that case, we reach the stable final position (1), characterized by
the van der Waals forces being very strong at such short distances. The dispersion is
now irreversibly coagulated and cannot be regenerated (repeptized) again because now
not only the energy between zero and primary maximum has to be applied, but also
between the primary minimum and maximum.
The particles lose their identity at this point and start to fuse, beginning the film
formation, ▶Section 7.9.2.

6.3.3 Stabilization of dispersions, dispersants

As we have seen before, pigment or polymer dispersions are only stable under certain
conditions. While coalescence and film formation are desirable in (1) after painting, the
system must be previously stabilized at (3) as a dispersion to avoid flocculation, that
means during storage and processing of the acrylic paints, [218, Chapters 2, 5, 9], [221,
Chapter 4], [185, Chapter 3.2, [Link]].
For stabilization, we need to prevent the reversion (3 → 1) to the thermodynam-
ically more stable flocculation state through kinetic barriers. We can choose between
several ways:
6.3 Wetting agents and dispersants, grinding paints � 553

– Electrostatic stabilization: We add electrolytes to the dispersion with a high affinity


for the polymer and surround it with a shell of similar electrical charges of the same
type. The charged polymers repel each other. Using this method, we increase the
barrier Em .
– Steric stabilization: We add protective colloids, which, section by section, alternately
have high polymer or water affinity. The hydrophilic sections are designed to be
spatially demanding. In this way, the polymers acquire a bulky or “hairy” surface to
keep the particles at a distance. Again, we increase Em by adding repulsion terms.
– Addition of surfactants (emulsifiers): We add these to reduce the surface tension
of the polymer particles significantly. The surface energy of the polymer particles
contributes to the total energy of the polymer system according to

UO = γ ⋅ A (6.13)

Herein, γ is the surface tension, and A is the surface area. The smaller these particles
are, the greater their surface area and surface energy. The system can reduce its
total energy by reducing the surface area and forming a few large particles from
many small ones in the dispersion. We perceive this as coagulation, flocculation,
phase separation, or precipitation of solids. In this process, the polymers approach
each other, and the van der Waals forces provide for further mutual adsorption of
the particles.
The surfactant significantly reduces the surface energy of the system after ▶equa-
tion (6.13) and also the possible energy gain in flocculation. Therefore, despite this
energy gain, the particle system cannot overcome the barrier Em on its own.

Electrostatic stabilization
The electrostatic stabilization of dispersions is based on the coverage of the polymer
particles with the same electrical charge by adding ionic compounds, ▶Figure 6.8. Sim-
ple compounds have a hydrophobic adhesive group and a hydrophilic end; they thus
correspond to the ionic surfactants we discussed in ▶Section 6.3.1 and show that these
dispersants can also act as wetting agents simultaneously. Examples of surfactants in-
clude alkyl polyethoxy sulfonates, alkyl polyethoxy sulfates, alkyl benzene sulfonates
with chain lengths n between 10–30, or polyphosphoric acid.
Polymer surfactants have several hydrophobic and hydrophilic sections. An exam-
ple is polyacrylic acid PAA, used as potassium, sodium, or ammonium salt, which obtains
hydrophilic and hydrophobic sections by copolymerization with acrylic acid esters. Such
polymers have a stabilizing effect partly also because of their steric demands.
In the case of polymer dispersions, it is possible to introduce charges by copolymer-
ization with suitably charged monomers in the polymer chain itself. One monomer is
acrylic acid, anionically charged in a suitable environment (acrylate anions).
The hydrophobic ends of the surfactants are adsorbed on the polymer (e. g., via
van der Waals interactions), while their hydrophilic (charged) ends protrude into the
554 � 6 Structure of paint systems

Figure 6.8: Electrostatic stabilization of polymer or pigment dispersions. The particles are covered with
ionic surfactants in a hedgehog-like manner. The symbol ⊖ symbolizes charged hydrophilic heads or chain
sections; zigzag sections symbolize hydrophobic hydrocarbon chains.

aqueous solvent and build up the charge layer. To compensate for the charge, the coun-
terions outside the charged polymer particles form an oppositely charged cloud so that
we have an electrically charged double layer with a diffuse flow to the outside.
The electrostatic repulsion drives similarly charged clouds, and thus also polymer
particles apart and stabilizes the dispersion in this way. Since external factors easily
influence the electrical conditions, we must not add electrolytes such as salts, acids, or
bases to such a stabilized dispersion without further consideration because the charge
distribution and layer structure would be changed or destroyed. The consequence
would be a coalescence of the dispersion.

Steric stabilization
This method of dispersion stabilization is based on coating the polymer with nonionic
surfactants or protective colloids, ▶Figure 6.9. Alkyl polyglycols are used as surfactants;
with a chain length n between 10–50, they are significantly larger than ionic surfac-
tants.
Again, the hydrophilic parts of these molecules show affinity to the aqueous phase,
while hydrophobic parts are oriented towards the polymer (e. g., due to van der Waals
interactions) and cover it densely. This protective layer is thin compared to the particle:
while pigment and latex particles have a size of about 0.1–10 µm, the protective layer is
about 20 nm thick. Here again, we can use simple or polymeric stabilizers. Two driving
forces are discussed as possible mechanisms for this type of stabilization; both prevent
the dispersed polymer particles from approaching each other too closely:
6.3 Wetting agents and dispersants, grinding paints � 555

Figure 6.9: Steric stabilization of polymer or pigment dispersions: The dispersed particles are covered
in a hedgehog-like manner with nonionic surfactants or protective colloids. The dark dots ∙ symbolize
hydrophilic heads or chain sections; zigzag sections symbolize hydrophobic hydrocarbon chains.

– When two particles approach each other, the layers of the stabilizers permeate,
which restricts the conformational mobility of the stabilizer molecules. Due to the
increasing order, the entropy decreases, and the system’s internal energy increases.
The level of this energy threshold is the barrier against aggregation of the polymer
particles.
– As the stabilizer layers permeate, the concentration of the stabilizer increases, and
solvent molecules are displaced from the overlapping space. If the stabilizing agent
has a high solubility in the solvent, solvent molecules try to dissolve it again to avoid
unfavorable stabilizer-stabilizer contacts and establish stabilizer-solvent contacts
instead. As a result, the solvent develops an osmotic pressure, which acts as a sepa-
rating force.

Protective colloids
In contrast to surfactants, protective colloids are high-molecular compounds but also
have hydrophilic and hydrophobic sections. Therefore, instead of many small surfactant
molecules, some of these large molecules of the protective colloid surround the polymer
particles. Examples include polyvinylalcohol (PVOH), poly-(vinyl-2-pyrrolidone) (PVP),
cellulose ethers such as hydroxyethyl cellulose (HEC), and polyacrylates such as poly-
methacrylic acid (pMAA), ▶Figure 6.10.
Depending on the ratio of the monomers, copolymers based on acrylic acid/acrylic
acid ester can be regarded as polymer surfactants or protective colloids. Spinelli [862]
shows by using the example of a dispersing aid for modern pigmented inkjet inks how
block copolymers p(A-B) assemble from hydrophobic (A) and hydrophilic (B) blocks. The
556 � 6 Structure of paint systems

Figure 6.10: Examples of protective colloids (nonionic compounds). Polymethacrylic acid exemplifies a
polymeric surfactant.

hydrophobic A block p(MMA/BMA/EHMA) contains methyl methacrylate (MMA), butyl


methacrylate (BMA), or ethylhexyl methacrylate (EHMA) as comonomers, which bind
to the hydrophobic pigment or particle. The hydrophilic B block is formed by includ-
ing acrylic acid (AA) or methacrylic acid (MAA) into a block of the same monomers,
B=A/AA/MAA. This hydrophilic region protudes into the aqueous phase.
Polyacrylates can be flexibly adapted through their composition and size of the A
and B blocks. For example, spatially demanding polymers with wide hydrophilic loops
for steric stabilization, or charged regions for ionic stabilization can form.

6.4 Thickener, rheology modifier


Thickeners and rheology modifiers adjust the viscosity and flow properties of paint
according to the application requirements. Typical agents are inorganic substances
(mainly silicates) or organic polymers (polyacrylates, polyvinylpyrrolidone, polyure-
thanes, and cellulose derivatives). Kittel [185, Chapter 4.3.1] extensively depicts rheology
in connection with paint, Hester and Squire [855] depict the thickening mechanism for
aqueous latex systems. Two processes achieve a thickening effect:
– Nonassociative thickeners or hydrogeling agents (▶Figure 6.11(a)) increase the vis-
cosity of aqueous systems. They have a thickening effect because they structure the
aqueous phase, and thus increase their hydrodynamic volume. The dipole moment
of the water molecules is exploited. Water molecules are adsorbed on the thick-
ener’s polar groups and also form hydrogen bonds. Above a critical concentration,
entanglement of polymer coils occur, enhancing the viscosity.
6.4 Thickener, rheology modifier � 557

Figure 6.11: Two ways of thickening a solution. White circles: polymer particles; black dots: pigments.

The resulting gels are often structurally viscous since the hydrogen bonds break
when the shear force is sufficiently large. As a result, they become highly fluid when
subjected to external forces (brushing, spraying, stirring) but regain their viscosity
as soon as the external force decreases. If this change in viscosity does not occur
immediately with the onset or end of force but rather with a delay, we speak of
thixotropic behavior.
– Associative thickeners (▶Figure 6.11(b)) are hydrophilic polymers with hydrophobic
subregions. In contrast to hydrogelling agents, they do not involve the solvent but
the other constituents of the paint dispersion, such as pigments or binder polymers.
While the thickener polymer disentangles in the water, the hydrophobic regions are
adsorbed onto other thickener polymers, pigments, and binder polymer particles,
thus forming a network that includes all components except water. The increase in
viscosity is caused by particle bridging.

Nonassociative thickeners (hydrogeling agents)


Nonassociative thickeners structure the aqueous phase, and thus increase their hy-
drodynamic volume, ▶Figure 6.11(a) [185, Chapter [Link].1.1]. At the same time, hydro-
colloids can act as protective colloids. Nonassociative thickeners are polymers of high
molecular weight. Examples are:
– cellulose ether
– other polysaccharides (starch, guaran, xanthan)
– homopolymers of unesterified acrylic acid p(AA)
558 � 6 Structure of paint systems

Cellulose-based nonassociative thickeners


Cellulose-based nonassociative thickeners are cellulose first converted into sodium
hydroxylate by boiling with sodium hydroxide solution. Subsequent treatment with
chloroethane, ethylene oxide, propylene oxide, or chloroacetic acid leads to ethers of
ethanol, propanediol, or acetic acid [185, Chapter [Link].1.1]. Etherification occurs pref-
erentially at the primary hydroxyl group C6 , followed by C2 and C3 . The hydroxyethyl
ethers formed in the reaction of ethylene oxide with primary hydroxyl groups can fur-
ther react to form polyethoxy ethers [856]. Commercially, cellulose ethers are present
in products such as Klucel® E or Tylose® :
– methyl cellulose (MC)
– methyl hydroxyethyl cellulose (MHEC, Tylose® MH [990])
– methyl hydroxypropyl cellulose (MHPC, Tylose® MO [990])
– hydroxyethyl cellulose (HEC, Tylose® H, HS, HA [990])
– hydroxypropyl cellulose (HPC, Klucel® [988])
– sodium carboxymethyl cellulose (NaCMC) or sodium carboxymethyl hydroxyethyl
cellulose (NaCMHEC)

As the formula structures show, only some hydroxyl groups of the cellulose are etheri-
fied, the exact number determining the properties of the individual products. Typically,
0.8–1.3 of the three free hydroxyl groups are etherified:
6.4 Thickener, rheology modifier � 559

The modifications of cellulose shown are necessary to convert it into an appropriate


water-soluble form. In its normal state, cellulose is insoluble in water because several
cellulose chains can be clustered together in a rope-like manner, ▶Section 6.7.1 at p. 569.
Hydroxyl groups reinforce the cohesion by forming numerous hydrogen bonds, giving
the strands high strength. Due to this ability to form microcrystalline adducts, cellulose
is a part of many natural structural materials such as wood, cell walls, or stale bread.
Etherification converts the water-insoluble cellulose into a soluble form as the ethers
block hydroxyl groups. This process prevents the formation of hydrogen bonds and the
clustering of strands into insoluble adducts.
Cellulose ethers increase the viscosity of the aqueous phase by adsorbing water
molecules along their chains. Thus, their hydrodynamic volume expands considerably.
The viscosity expands with the molecular weight of the cellulose ethers and concen-
tration and falls with temperature. Cellulose tangles can be disentangled by high shear
forces and aligned along the direction of the force so that cellulose-thickened solutions
are strongly shear thinning. As a result, they become more fluid when subjected to ex-
ternal forces, such as during painting, rolling, spraying, or stirring. High degrees of poly-
merization enhance this property.

Starch
Starch paste also serves as a thickener [185, Chapter [Link].1.1.2]. Since amylose tends to
form crystallites on cooling, starch is modified with ethylene oxide or propylene oxide.
The resulting hydroxyethyl starch (HES) or hydroxypropyl starch (HPS) is readily water-
soluble and stable.
560 � 6 Structure of paint systems

Polyacrylic acid-based nonassociative thickeners


Homopolymers of free acrylic acid (PAA) or methacrylic acid (PMAA) can also act
as nonassociative thickeners [185, Chapter [Link].1.2]. Polyacrylic acid adsorbs water
molecules along its chain and, in this way, increases its hydrodynamic volume. By
copolymerization p(AA/MAA-EA) of polyacrylic acid with acrylic acid esters such as
ethyl acrylate (EA), the polyacrylic acid is partially hydrophobically modified, and this
controls the extent to which ambient water is structured.

ASE 60® is a commercial, partially hydrophobically modified polyacrylic acid based on


a copolymer of methacrylic acid with ethyl acrylate p(MAA-EA), ▶Figure 6.12 [987]. ASE
stands for alkali soluble emulsion and is an acid, thin emulsion of undissociated poly-
acrylic acid in the delivery form. It reacts during neutralization with a substantial in-
crease in viscosity. Carboxylic acid groups are transformed into carboxylate groups. The
resulting homogeneous charge causes a disentangling of the acrylic polymer, which ad-
sorbs water via hydrogen bonds in a wide area, thus forming an extended wide-meshed
gel. Therefore, polyacrylate thickeners are strongly pH-dependent and exhibit a maxi-
mum effect at medium pH values.

Figure 6.12: Swelling of nonassociative thickeners.


6.4 Thickener, rheology modifier � 561

Neutralizing agents for polyacrylic acid are salts of strong bases (alkali hydroxides, am-
monium hydroxides, or alkali carbonates) or organic amines (triethanolamine, diiso-
propylamine, 2-amino-2-methylpropanol-1) [986]:

Ammonium salts or amines with high vapor pressure are preferable to alkali salts since
they can evaporate during the film drying; they partially evaporate while the free car-
boxylic acid reforms. On the other hand, alkali anions remain permanently in the film
and lead to constant redissolving.

Cross-linked polyacrylic acid thickener


ASE thickeners can cross-link with bifunctional dienes, an example being the Carbopol®
EZ-2, frequently available in artists’ supplies. It is polyacrylic acid (PAA) with small parts
of acrylate components, which is cross-linked with polyalkenyl ether or divinyl glycol
(hexa-1,5-diene-3,4-diol), ▶Figure 6.13 [986].

Figure 6.13: Cross-linked polyacrylic acid (PAA). Thick lines: PAA chains connect through bridges. Thin
lines: PAA chains form large-scale networks. The carboxylate groups structure the water through hydrogen
bonds (hydrogel formation). White circles: other components of the paint (pigments or particles of a latex
dispersion), which are adsorbed on the polyacrylate polymers or held in the meshes of the network.

Cross-linking of the polyacrylic acid chains is achieved by replacing one acrylic acid unit
with a bifunctional diene that can become part of two chains and links them:
562 � 6 Structure of paint systems

Other cross-linking monomers include trimethylolpropane diallyl ether (TMPDAE),


dipropyleneglycol diacrylate (DPGDA), and allyl methacrylate:

Due to the cross-linking, the particles are no longer physically soluble but only col-
loidally. They can swell considerably, carbopoles to a thousand times their original vol-
ume. A high water-binding capacity accompanies this increase, so we often find cross-
linked polymers in products such as diapers or absorbent gritting material.

Associative thickeners
Associative thickeners are water-soluble polymers of low molecular weight, but with
numerous hydrophilic groups and distinct hydrophobic regions [185, Chapter [Link].1.3].
They thus resemble polymer surfactants used to stabilize a dispersion. Their hydrophilic
sections extend into the aqueous phase, and hydrophobic sections become adsorbed on
pigments or polymer particles, ▶Figure 6.11, right. Gel formation occurs through hy-
drophobic interaction between alkyl chains, leading to increased effective molecular
mass and a wide-meshed network. In contrast to hydrogel-forming agents, this gel does
not involve the solvent, but the other particles of the ink, such as pigments or dispersed
binder polymers.
To obtain associative thickeners, we can copolymerize numerous hydrophilic poly-
mers that serve as thickeners with hydrophobic monomers. Commercially available
products on the market can be identified by the designation “hydrophobically modi-
fied”: ether-urethanes (▶ HEUR), polyethers (▶ HMPE), polyacrylamides (▶ HMPAM),
polyacrylic acids/ASE (▶ HASE), and hydroxyethyl cellulose (▶ HMHEC).
Hydrophobic elements are introduced by incorporating small amounts of “hy-
drophobically modified monomers” (HMM) at both ends of the polymer chain or ran-
domly scattered along its length. Usually, alkyl polyglycols, alkyl aryl polyglycols, or
6.4 Thickener, rheology modifier � 563

fatty alcohols are hydrophobic carriers introduced as acrylate or methacrylate into a


polyacrylic acid, or they form cellulose ethers with cellulose derivatives. Exemplary
monomers are C16−22 -alkylated compounds [856, 858, 859, 861, 863–865, 989]:

HASE thickener
Hydrophobically modified ASE thickeners (HASE) are copolymers p(AA/MAA/MS/Ita-
conic acid-EA-HM) of acrylic or methacrylic acid (ASE) with a hydrophobic, long-chain
alkyl comonomer (HM) and possibly further acrylic acid components such as ethyl acry-
late, butyl acrylate, or methyl methacrylate [185, ch. [Link].1.3.4], [856]. HM is an alkyl
polyglycol or alkyl aryl polyglycol that is introduced as a (meth)acrylate or maleinat
into the chain, which may also contain maleic acid or itaconic acid:

An example from commerce is Tafigel® AP with the structure p(AA/MAA-HM), where


HM denotes the typical hydrophobic monomer alkyl-polyethoxy-acrylic acid ester [989].
The proportion of hydrophobic comonomers is about 20 %.
564 � 6 Structure of paint systems

Due to the composition, HASE thickeners have a comb-like shape, the hydrophobic
chains forming the teeth, ▶Figure 6.14(a). Like ASE, the hydrophobically modified prod-
uct must be neutralized with alkaline solutions to initiate the swelling process. Besides
the associated electrostatic repulsion, the swelling is supported by the fact that the glass
temperature of the polymer drops sharply in contact with water. This drop consider-
ably increases the mobility and radius of movement of the polymer chains, which is
perceived macroscopically as swelling.

Figure 6.14: Structures of associative thickeners used in paints.

HEUR thickener
Hydrophobic polyurethanes are also used in water-based paints [185, Chapter
[Link].1.3.1]. This HEUR type includes products such as Tafigel® PUR [989]. Their struc-
ture is an ABA block copolymer with polyethylene glycol as the B block, sectionally
linked to hydrophobic urethane units (A block), ▶Figure 6.14(b). A long-chain hydroxyl
or amino compound is chosen as the alkyl chain, as the following example shows [866]:
6.5 Film-forming aids (coalescing agents) � 565

HM*C thickener
If we provide cellulose ethers with hydrophobic units, we arrive at HM*C types [185,
Chapter [Link].3.5], [856]. An example is HMHEC hydrophobically modified with ran-
domly distributed alkyl chains. Typically, 5 % of the hydroxide groups are alkylated, and
the chain length of the alkyl residues varies between 10–24:

6.5 Film-forming aids (coalescing agents)


Film-forming aids or coalescing agents intend to optimize the film formation of binders
based on a polymer dispersion under specific application conditions or even make them
possible in the first place [190, Chapter 6.2].
If a polymer dispersion is used as a binder, as with acrylic paints, a perfect film
formation is only possible above a specific minimum temperature, the so-called mini-
mum film-forming temperature, MFT. The last step of film formation is the interdiffusion
(depicted in more detail in ▶Section 7.9.2), in which polymer particles are welded to-
gether by the exchange of polymers, thus forming the homogeneous film. This phase
takes place after drying. Toward the end of the actual drying, capillary forces increas-
ingly occur, which are the more significant, the thinner the remaining water films be-
tween the polymer particles become. These forces press the polymer spheres against
each other, whereby the spheres must be increasingly hexagonally deformed. In addi-
tion, by fusing individual particles to form larger units, the particles can reduce their
surface energy, as the surface area of the fused units is smaller than that of the numer-
ous individual particles.
The hardness or deformability of the polymer spheres influenced by the tempera-
ture during filming opposes this endeavor. Therefore, filming is only possible at or above
the MFT; only then do the polymer particles become sufficiently deformable to coalesce.
The polymer’s MFT and glass transition temperature TG are closely related; hard poly-
mers lead to a high MFT, soft ones to a lower one. For standard artist acrylic dispersions,
e. g., the MFT is approximately at room temperature or slightly below.
In order to allow processing at lower temperatures, soft polymers with low TG can
be applied, but these result in soft films and are therefore not suitable for the desired
purpose. Film-forming aids temporarily lower the MFT to such an extent that a uniform
film formation is possible even at lower temperatures. They are initially dissolved in the
566 � 6 Structure of paint systems

aqueous phase and diffuse into the polymer particles, reducing their hardness and in-
creasing their deformability, i. e., act as plasticizers. The film-forming aids gradually es-
cape from the film during drying. The film hardens as the agent evaporates and regains
its original hardness determined by the TG . This temporary softening allows harder poly-
mers, which produce a film with high abrasion fastness, a more uniform surface, and
consistent color reproduction due to the improved filming.
These auxiliaries, by nature, are moderately hydrophilic solvents that evaporate
more slowly than water and must have a particular affinity for both water and polymer.
Therefore, high-boiling polar products with alcohol, ester, ether, or keto functions are
preferred. In addition, frequently used are esters of lower alkane carboxylic acids, and
alkyl ethers of ethylene glycols and propylene glycols, ▶Figure 6.15.

Figure 6.15: Examples of film-forming aids (coalescing agents) commonly used in paints [190, Chapter 6.2].

6.6 Other excipients


Defoamers prevent foaming during processing and support air escape during the drying
of the binder film. Fatty acid esters, metal soaps, mineral oils, silicone oils, and siloxanes
are used.
Siccatives support the drying process of a chemically drying binder and promote
its radical polymerization. Metal soaps with cobalt, manganese, calcium, zinc, or bar-
ium are used. In ▶Section 7.4.5 at p. 675, we learn about applying a siccative using the
example of oil painting.
Preservatives prevent infestation of the moist and often nutrient-rich colors
(binders!) by microorganisms.
6.7 Paper � 567

6.7 Paper
As one of the earliest artificial writing materials for texts and paintings, papyrus was
used already 3000 BC in Egypt [777]. Paper has not such a long tradition; depending
on the continent, it is longer (Asia) or shorter (Europe). Paper as we know it was in-
vented in China in the year 105; in Europe, we find papermills from 1144 (Spain) and
1390 (Nuremberg) onward. The fundamental manufacturing processes and materials
have changed little for an extended period. However, in the last few decades, chemistry
has produced numerous specialty papers through selective surface coatings and paper
pulp treatment, including writing paper to filter papers to absorbent, wet-tear resistant,
or oil-resistant paper. At first glance, these finishing techniques seem unspectacular for
the chemistry of painting, but they prove to be a treasure trove [820–822, 824, 825], [218,
p. 301], [1005]. [173–178], [203, keyword “paper”], [182, 183, 830] consider the production,
properties, and chemistry of paper in detail.
Chemically, the paper consists of densely matted fibers and can form sheets of var-
ious thicknesses. Depending on the raw materials the fibers are extracted from, we ob-
tain high-quality papers for writing and drawing, cheap ones for mass printing, card-
board papers, or high-quality rag paper. The various products differ in the type of fibers
and the number of manufacturing and finishing steps. ▶Figure 6.16 illustrates the es-
sential steps of paper production:
– Extraction of the fibrous materials from the raw materials, e. g., of pulp (lignin-free
cellulose) from wood, or conversion of the raw materials into a suitable fibrous
form.
– Production of the paper pulp from the fibrous materials to scoop the paper. As a rule,
a mixture of different fibrous materials is used to obtain the desired paper grade.
– Addition of additives to control dry and wet strength, fixatives to fix certain chem-
icals on the paper, and flocculants to facilitate processing.
– Addition of special additives to control surface tension (sizing in the pulp) and
colorize (coloring in the mass).
– Scooping (traditionally by hand) or screening (by machine) the sheets from the pulp
to produce the paper web. By pressing off the water with felts, the surface gains a
texture (rough, matte, ribbed) and, optionally, a watermark.
– If necessary, surface treatment for surface coloring of the paper and control of the
parameters smoothness and absorbency.
– If necessary, calendering (rolling) to achieve smoother, shimmering surface quali-
ties.

6.7.1 Structure and composition of raw materials

Paper-building fibers can come from a variety of sources. In principle, using many raw
materials such as straw would be possible. However, in the countries of origin of the pa-
per, papyrus plants were a natural choice. In the European region, cellulose fibers from
568 � 6 Structure of paint systems

Figure 6.16: Overview of the main raw materials for paper fibers and the main processing steps during
paper production [173–178], [203, keyword “paper”], [830].

rags, or textile cotton waste, established themselves as a raw material for centuries.
However, woods replaced them as the primary paper source in the last two centuries.
Today, synthetic fibers are the choice for tear- and water-resistant specialty papers. For
high-quality papers, however, rags are still used today. Recycled papers contain recycled
paper products such as market deinked print (MDIP) or deinked pulp (DIP, recovered pa-
per with ink removed), recycled bleached kraft (RBK, recycled bleached Kraft paper), old
corrugated container (OCC, used cardboard boxes), or old newspaper (ONP). However,
such raw materials are not considered for long-life artists’ papers.
The selection of raw materials and processing methods before preparing the pulp
determines the essential parameters of the paper. The most important raw materials
are woods, cotton, and synthetic fibers; in the Middle Ages, rags from textiles containing
cellulose. Approximately 43 of the pulp needed today is produced chemically, 41 by me-
chanical means. Of the chemical pulps, 90 % are produced by the Kraft process, 10 % by
6.7 Paper � 569

the sulfite process. Cotton fiber pulp is indispensable for producing high-quality artists’
papers (and banknotes) but is insignificant in quantity.
The most critical target parameters for paper are strength and printability. Soft-
woods (spruce, fir) yield longer fibers (2.5–4.5 mm) than hardwoods (beech, birch, oak,
0.7–1.6 mm) and give the paper higher strength, whereas fillers reduce it. Conversely, the
printability of hardwood pulp is better because the short wood fibers compensate for
surface irregularities more adequately than long fibers. Specific processing steps such
as sizing, coating, and calendering improve the surface properties.

Wood
The most widely available renewable raw material, wood, consists of approximately
40–45 % cellulose, 25–35 % hemicelluloses, 20–30 % lignin, and small amounts of other
substances such as waxes or fats, depending on the type of wood, [173], ▶Table 6.4. The
most valuable part of the wood includes cellulose and hemicelluloses; in the case of
cheap papers, lignin is also essential for increasing the paper mass. However, lignin de-
grades over time under the influence of light to various colorless and yellowish com-
pounds, leading to the well-known rapid yellowing of wood-containing papers, ▶Sec-
tion 6.7.8.

Cellulose
The essential component of paper fiber is cellulose, a chain-like molecule of poly-(β-
D-1,4-glucopyranosyl-glucopyranoside) [173]:

In cotton fibers, the chains contain about 1500 glucose units; in spruce wood, about 1500;
and in processed pulp, about 700.
The β linkage of the glucose molecules results in a complete turnaround of the chain
itself every two glucose units. This glucose pair is called cellobiose:

Specific functional groups of cellulose are significant for paper chemistry and dyeing:
– Per glucose unit, one primary and two secondary alcoholic hydroxyl groups are
contained. These groups are responsible for the main properties of cellulose in pa-
570 � 6 Structure of paint systems

per chemistry, as they can bridge several cellulose chains or fibers via hydrogen
bonds.
– Carboxyl groups replace approximately 1 % of all primary hydroxyl groups due to
growth-related irregularities. This small proportion of carboxyl groups is also vital
for paper chemistry, as it allows cellulose to dissociate in the aqueous medium.
The resulting polycarboxylate exhibits anion activity and influences the binding of
auxiliaries, sizing and coating, and dye affinity.
– Aldehyde, keto, or (glycosidic) hydroxyl groups can occur per terminal glucose unit
instead of the standard hydroxyl groups. Cleavage of the terminal glucopyranose
ring is also possible:

The alcoholic hydroxyl groups and the ring oxygen cause the high strength that cellu-
lose fibers can develop: as soon as cellulose strands are oriented in parallel, they can
stiffen considerably due to numerous hydrogen bridges, as ▶Figure 6.17 shows schemat-
ically [280, Chapter 4.4.16]. Hydrogen bridging is possible only in quasicrystalline, i. e.,
ordered regions of cellulose since a distance of the binding partners of about 0.25 nm is
necessary for forming the bridge. In amorphous (disordered) regions, the chains are too
far away from each other, and the density of the hydrogen bonds is considerably lower.
In the same way, the cellulose chains on the surface of the fibers are bridged. Con-
sequently, they form the desired paper felt in the pulp during the paper sheet’s scoop-
ing process. The hydrogen bonds are thus responsible for the dry strength of the paper.
A higher paper strength requires denser bridging, longer fibers, or the addition of syn-
thetic polymers that act as bridges (dry-strength additives).
The mechanism explains the considerably lower wet strength of untreated paper
compared to the dry state: Water molecules move between the cellulose chains and in-
terrupt the direct bridging of the chains. As a result, they act as a gliding layer, and the
close cohesion of chains and fibers is lost, thereby most of the strength. Therefore, if
desired, wet strength must be ensured by adding cross-linking auxiliaries (wet-strength
additives).
Water ingress is particularly possible in amorphous areas, which can be exploited
when producing specific water-soluble cellulose derivatives that act as glues and rhe-
ology modifiers. Sterically demanding substituents keep the cellulose chains artificially
6.7 Paper � 571

Figure 6.17: Schematic representation of the bridging of two cellulose chains via hydrogen bonds. Left: in
the dry state. Right: in the wet state. Blue: In the latter state, water molecules approach, push between the
chains and disconnect them. In this way, the number of connections between cellulose chains or fibers,
and thus the strength decreases.

separated from each other and prevent the formation of crystalline regions. The sub-
stituents are introduced by methylation or esterification of hydroxyl groups, ▶Sec-
tion 6.4 at p. 558.
The aging processes of cellulose, which lead to the decay of the paper and occur
particularly in the acidic environment, will be discussed in ▶Section 6.7.8.

Hemicelluloses
Hemicelluloses are heterogeneous polysaccharides consisting of up to 200 monosaccha-
rides linked by 1 → 4- and 1 → 3-glycosidic bonds and branched to a small extent, ▶Ta-
ble 6.2 [173, Chapter 5], [176, Chapter 7.4], [268, 269]. The hexoses glucose (Glc), mannose
(Man), and galactose (Gal), as well as the pentoses xylose (Xyl) and arabinose (Ara), ap-
pear as monomers. As functional groups, they possess primary and secondary hydroxyl
groups, which are practically not dissociated but influence the properties of the hemicel-
luloses by forming hydrogen bridges. In xylans, the uronic acids glucuronic acid (GlcA),
galacturonic acid (GalA), and 4-O-methyl-glucuronic acid (4-OMe-GlcA) appear, which
are largely dissociated by their pKa of 4–5 at a pH value of 7 and contribute strongly to
the anion activity of pulp.
The linear glucomannans consist of the main chain of (1 → 4)-linked β-D-Manp and
β-D-Glcp and carry side chains of β-D-Xylp or acetyl groups:
572 � 6 Structure of paint systems

Table 6.2: The main hemicelluloses in wood and their composition [173, Chapter 5].

Wood type Hemicellulose Proportion % Monomer Molar Bond


Proportion

Softwood Galactoglucomannan 5–8 β-D-Manp 3–4 �→�


β-D-Glcp 1 �→�
α-D-Galp 1 �→�
O-Acetyl 1
Softwood Glucomannan 10–15 β-D-Manp 3–4 �→�
β-D-Glcp 1 �→�
α-D-Galp 0.1 �→�
O-Acetyl 1
Softwood Arabinoglucuronoxylan 7–15 β-D-Xylp 10 �→�
4-OMe-α-D-GlcpA 2 �→�
α-L-Araf 1.3 �→�
Larch Arabinogalactan 3–35 β-D-Galp 6 � → �, � → �

L-Araf �
�→�

β-D-Arap �
�→�
Hardwood Glucuronoxylan 15–35 β-D-Xylp 10 �→�
4-OMe-α-D-GlcpA 1 �→�
O-Acetyl 7
Hardwood Glucomannan 2–5 β-D-Manp 1–2 �→�
β-D-Glcp 1 �→�
O-Acetyl 1

The highly branched galactans have the main chain of (1 → 3)-linked β-D-Galp and side
chains of β-D-Galp and α-L-Araf :

The acidic xylans possess a backbone of (1 → 4)-linked β-D-Xylp monomers and carry
short side chains of 4-OMe-α-D-GlcpA, α-L-Araf or acetyl groups. They contribute deci-
sively to the anion activity of pulp through their carboxyl groups:
6.7 Paper � 573

Hemicelluloses are present in large quantities in raw or processed wood and are the
primary source of the anion activity of the paper pulp, and thus responsible for techno-
logical effects, among others:
– They influence the water absorption and swelling behavior of the paper.
– They bind auxiliaries. Essential for us is the binding of cationic dyes.
– They form a swollen viscous mucilage, which acts as an adhesive and considerably
strengthens the fiber mass during drying.

Hemicelluloses are among the valuable constituents preserved during wood process-
ing whenever possible since their content of uronic acids makes them anionic polyelec-
trolytes. Unfortunately, since they are degraded via hydrolytic 1 → 4 cleavages more
rapidly than cellulose, they are also one of the causes of paper aging.

Lignin
Lignin is a hard, highly polymeric phenolic substance that, as an extensive network,
fills large parts of the wood structure as a fundamental matrix. Carbohydrates (cel-
lulose, hemicelluloses) are embedded in this matrix. Wood gains structural strength
from the compression-resistant lignin, while the tear-resistant and flexible carbohy-
drate fibers contribute to the high tensile strength like a concrete reinforcement. Basic
building blocks of the high polymer amorphous polycondensate are C3 −C6 elements,
so-called phenylpropanes [173, 762, 763], [275, Chapter 2.1]. The main building blocks
are

In lignin, these building blocks are linked in various ways through carbon and ether
bonds. The formula shows a possible section:
574 � 6 Structure of paint systems

It contains typical elements such as ether bridges, alcoholic and phenolic hydroxyl
groups, and aryl-α-carbonyl structures, which are essential for the chemistry of the
pulp and the finished paper.
The aim of obtaining the so-called chemical pulp is to reduce the lignin content
as far as possible since lignin forms complex colored quinoid compounds through ox-
idation in air. These are the primary source of yellowing phenomena in cheap wood-
containing (meaning lignin-containing) papers, ▶Section 6.7.10. As a hydrophobic and
water-insoluble compound, lignin also prevents the formation of hydrogen bonds be-
tween cellulose fibers, reducing the paper’s strength.

Color of wood
The question of the nature of wood color is of practical importance since it directly influ-
ences the color of the pulp [174, Chapter 3.3, 3.4]. Wood has a yellow-brown color from
the beginning, caused by specific chromophoric structures within the lignin network.
In addition, wood, bark, and pulp can also darken by phenolic structures oxidatively
reacting with oxygen to form quinones or with metal cations to form metal complexes.
The primary chromophore in wood is coniferylaldehyde, which occurs at a fre-
quency of 5 per 100 phenylpropane units in lignin. Due to the absorption maximum
of 400 nm in the solid form, this compound absorbs blue light and imparts a yellowish
color to the wood. α-keto structures, o- and p-quinones, and dienones also contribute to
the color.
6.7 Paper � 575

Free phenolic hydroxyl groups, traces of o- and p-quinones, and catechol structures can
be converted simply to the corresponding quinones if oxygen is available (autoxidation).
Typical structures include:

Quinones react with oxygen and water in a neutral or alkaline solution to chromophores
that are more intensely colored. With other phenols, e. g., from water-soluble bark sub-
stances, further intensely colored structures can be obtained via autoxidation and phe-
nol condensation, ▶Figure 6.18. At temperatures of 140–170 ℃ maintained for TMP pro-
duction, even more chromophores with conjugated carbonyl and double bond struc-
tures form.

Figure 6.18: Formation of intensively colored structures from wood in an alkaline environment by autoxi-
dation and phenol condensation [174, Chapter 3.4].

Cotton, linters
Cotton fibers are seed hairs of certain members of the family malvaceae, i. e., the mal-
low family [184, Chapter 9]. The fibers used for textiles come from the highly visible
white fiber tuft surrounding the mature seeds. Depending on the plant, these fibers are
22–36 mm in length and are readily spinnable to produce yarns and, eventually, fabrics.
Each cotton fiber represents a single, long cell that develops in the surface layers
of the boll. The matured fiber is dead and consists of a hollow, twisted cell wall tube.
576 � 6 Structure of paint systems

The length of the fiber and its twists give cotton the typical grip that distinguishes it
from other fibers. Chemically, cotton consists of > 95 % cellulose; the cotton fibers con-
verge into polymer tangles, giving them high strength. This property also makes them
extremely valuable as a raw material for artists’ papers where tear-resistance and dura-
bility are required.
For artists’ papers, however, not only the long cotton fibers of the textile industry
are eminent, but also a second type of fiber that grows inside the white fiber tuft directly
at the boll tip. These are called linters. Their fibers are very short (12–15 mm), sometimes
pigmented light brown, and have different chemical and physical properties. Thus, they
are coarser, stiffer, and about twice as thick as textile cotton fibers. Consequently, they
are not suitable for textile production but indispensable for pulp production.

6.7.2 Pulp from wood

As shown above, paper can be made from various raw materials. The raw materials are
first turned into a mash called the pulp, which shows typical properties according to
the raw material and the manufacturing process. Then, depending on the desired paper
grade, the pulp is processed into paper or blended with other pulp types.
In many cases, the choice falls on wood as the cheap sole or primary raw material of
the paper. After debarking the logs, there are two basic methods for processing the logs
into pulp, the mechanical comminution of the entire wood log or the chemical isolation
of the cellulose. In both cases, recovering the cellulose fibers contained in the wood is
the objective of pulp production. The pulp exhibits specific properties characteristic of
the production method, ▶Table 6.3.

Table 6.3: General properties of mechanical and chemical pulp [174].

Mechanical pulp (MP, wood pulp) Chemical pulp (CP, cellulose pulp)

High energy input No external energy required


90–98 % yield based on wood ≈ �� % yield based on wood, approx. 50 %
dissolved content
Contains all wood constituents except slightly Contains essentially cellulose, lignin is greatly
soluble carbohydrates reduced
Fibers are stiff and unchanged (▶ body, volume, Fibers are more flexible and can form
bending stiffness) quasi-crystalline regions (▶ strength)
Contains fibers, fiber fragments, and cell wall Contains essentially cellulose fibers
fragments (fines)
Less light in color, but opaque and smooth Light in color, solid
6.7 Paper � 577

[Link] Mechanical pulp, wood pulp


In mechanical pulp (MP) production, mechanical energy separates the cellulose fiber
bundles, fibers, and fiber fragments from the wood matrix (lignin) and the cell walls [174].
The focus here is on preserving the amount of wood pulp to a large extent to increase the
yield of paper pulp with adequate strength and brightness of the mash. The mechanical
comminution can be done by grinding the logs or by refining, ▶Figure 6.19.

Grinding wood, GW pulp


By grinding the logs, we obtain wood pulp of the type ground wood (GW/GWD) and stone
ground wood (SGW). During grinding, cellulose fibers and wood cell structures are torn
off as a whole or as fragments. The result is a high proportion of fine fiber and cell wall
fragments, the so-called fines. Pressure ground wood pulp (PGW) is produced similarly,
but the logs are preheated at 100–140 ℃ and then pressurized ground. The grinding pro-
cess is suitable for softwoods (spruce, fir) and hardwoods (beech, aspen).

Refining wood, RMP wood pulp


More gentle than grinding is crushing the wood into chips and separating the fibers by
refining between roughened discs. This so-called refiner mechanical pulp (RMP) contains
longer fibers, increasing the paper pulp’s strength significantly when the applied force is
adjusted to comminute the wood gently. The thermo-mechanical pulp (TMP) is cautiously
produced by treating the chips with hot steam at 115–155 ℃ and refining under pressure.
Chemi-thermomechanical pulp (CTMP) is produced even more gently. Here, the chips are
additionally impregnated with diluted sodium sulfite solution to dissolve the lignin and
facilitate the release of the long fibers. Softwood is treated with 1–5 % Na2 SO3 solution at
120–135 ℃, hardwood with up to 3 % Na2 SO3 - and 1–7 % NaOH solution at 60–120 ℃. The
chemical treatment reduces the yield by dissolving wood substances, especially resins.

Other types, effects on fiber content


Other pulp types differ from TMP and CTMP in applying pressure at the processing
stages and combining heat and chemical treatment. Thermo-refiner mechanical pulp
(TRMP) softens wood chips only with hot steam. Pressure refiner mechanical pulp (PRMP)
is comminuted under increased pressure; as a result, it exhibits increased strength. In
chemical mechanical pulp (CMP), the softening of the wood pulp takes place through
higher concentrated sodium sulfite (5–10 %), and the optical properties deteriorate
strongly. Chemical refiner mechanical pulp (CRMP) applies only mild chemicals for soft-
ening.
In all cases, the thermal or chemical pretreatment influences the fracture behavior
of the fibers during grinding or refining. With pretreatment or increased temperature,
the fracture zone shifts to the lignin-rich zone of the middle lamellae, especially when
578 � 6 Structure of paint systems

Figure 6.19: Main steps in the production of mechanical wood pulp and its common types (bold: most
common types) [830], [174, Chapters 2–4, 10].
6.7 Paper � 579

the temperature becomes higher than the softening temperature of the lignin (for soft-
wood such as spruce, about 125–145 ℃, for hardwood, about 20 ℃ less). As a result,
fewer fines and more long fibers are produced, increasing the strength at the expense
of opacity.
The fines are considered an advantage of mechanical pulp because they are respon-
sible for the opacity of the paper pulp and their good surface properties. The widespread
size distribution of the fines causes large fluctuations of the refractive index and high
scattering of light, i. e., high opacity. The high fiber stiffness in the mechanical pulp pro-
vides papers of low density or high volume, body, and high bending stiffness. If many
stiff, unmodified fibers are present, many fines must be available to act as a binder vir-
tually. A significant disadvantage of fines is that the lignin substances make the paper
susceptible to light and aging. In practice, the paper composition depends on the appli-
cation. Mechanical or chemical pulp or both are selected to combine the advantages of
both pulp types.

Impregnation/sulfonation
In the production of CTMP, wood is treated at elevated temperature with a 1–5 % solu-
tion of sodium sulfite Na2 SO3 , which dissolves resin out of the wood. Consequently, by
softening the lignin structure, easier separation of the fibers is achieved. The reaction
proceeds at a neutral or slightly alkaline pH value within minutes [174, Chapter 3.5]. With
a higher concentrated sulfite solution (5–10 %), chemimechanical pulp (CMP) is formed.
As a strong nucleophile, the sulfite anion attacks α,β-unsaturated carbonyl struc-
tures in lignin, as they are present in coniferylaldehyde or o- and p-quinones. As a result,
sulfonic acid groups add to the double bond or the quinone cores, increasing the water
solubility of the lignin, ▶Figure 6.20.

Figure 6.20: Addition of the strong nucleophile SO2⊖


3 to α,β-unsaturated carbonyl structures and o- or
p-quinones under the formation of sulfonic acids [174, Chapter 3.5]. The sulfonic acids increase the water
solubility of lignin.
580 � 6 Structure of paint systems

Bleach
The last step of pulp preparation is bleaching. In contrast to chemical bleaching, me-
chanical pulp bleaching must be carried out with the greatest possible care to preserve
the lignin substances. This bleaching aims to convert chromophores in the wood into
colorless forms, either to achieve uniform brightness or to increase the overall bright-
ness. The brightness of the mechanical pulp is initially 60–63 % ISO; by bleaching, up
to 80 % ISO can be achieved. The bleaching process is reversible, i. e., after some time,
a coloration may occur again. Therefore, mechanical pulps are intended for applications
more in the newspaper and magazine sector, not for fine office or even artists’ papers.
As a result, wood pulp grades are available such as bleached CTMP (BCTMP), special
bleached CTMP softwood (BCTMPSW), or bleached CTMP hardwood (BCTMPHW).
Mechanical pulp bleaching is carried out reductively with sodium dithionite Na2 S2 O4
in a neutral environment or oxidatively with alkaline hydrogen peroxide H2 O2 [174,
Chapter 3.6]. If the aim is only a uniform brightening of the pulp with dithionite, the
reagent can be added to the pulp. When bleaching with peroxide, the reaction is carried
out at 60–70 ℃ for 1–2 h in a separate stage.
In a bleach with dithionite, only certain structures such as o- and p-quinones are re-
duced, ▶Figure 6.21. Therefore, only partial bleaching occurs, and a yellowish coloration
remains.

Figure 6.21: Bleaching of mechanical pulp using dithionite S2 O2⊖


4 [174, Chapter 3.6]. Dithionite can only
reduce o- and p-quinones to colorless compounds.

The bleach with alkaline peroxide, on the other hand, allows a very high degree of bleach-
ing. The reaction must be carried out carefully because the pulp reacts readily to dark
substances by phenol oxidation in the alkaline environment. HO⊖2 is a strong nucleophile
and attacks electron-deficient structures such as coniferylaldehyde and conjugated car-
bonyl compounds. The reaction breaks down these structures, leading to smaller aro-
matic aldehydes and formic acid, ▶Figure 6.22, top. Hydroquinones can form if free hy-
droxyl groups are present, ▶Figure 6.22, bottom.
6.7 Paper � 581

Figure 6.22: Reactions in bleaching with alkaline peroxide [174, Chapter 3.6]. The strong nucleophile HO⊖2
attacks phenylogue double bonds and conjugated carbonyl structures. As a result, the lignin network
breaks, and aldehydes and small molecules such as formic acid emerge; in either case, the chromophoric
structure is destroyed. If free phenolic hydroxyl groups are present, hydroquinones can also occur.

The chromophores degrade in this type of bleaching. For example, the color-intensive
quinones are degraded to carbonyl acids but can form colored hydroxylquinones in a
competitive reaction toward the end of the bleaching reaction, when the concentration
of the hydroperoxide ions HO⊖2 is low:

[Link] Chemical pulp


The objective of chemical pulp (CP) production is to release the undamaged cellulose
fibers from the wood’s lignin matrix and remove the lignin to avoid aging and yellowing
effects. The fibers obtained by this method are more flexible than those of mechanical
pulp, and the fibers can adhere well to each other, thus giving the pulp a high strength.
Chemical pulp is, therefore, suitable for thick paper, sack paper, and cardboard paper.
Pulp is the substance remaining when all compounds except cellulose and hemicel-
luloses are removed from the wood, especially lignin. In practice, the cellulose content of
pulp is 60–85 %, besides a small amount of hemicelluloses and fiber wall lignin. Depend-
ing on the wood used (softwood, hardwood), the growing region (northern, southern),
and the type of bleaching, different types of pulp are produced and sold under individ-
ual names by pulp manufacturers, ▶Figure 6.23.
582 � 6 Structure of paint systems

Figure 6.23: The main steps in the production of chemical pulp (cellulose) from wood, as well as the lead-
ing types of pulp traded worldwide [830], [174, Chapters 5, 6]. Bold: most common types.
6.7 Paper � 583

The type of wood (softwood, hardwood) and the region in which the trees grow
(north, south) significantly influence the properties, as they determine the exact mor-
phology of the wood cells, and thus the cellulose fibers. Softwood yields tear-resistant
long fibers that assemble well to form a uniform surface or dense quasicrystalline ar-
eas. Therefore, softwood pulp such as NBSK is suitable as a reinforcing pulp that can
be added to a pulp mixture to give it strength. Papers become pliable and soft but tear-
resistant, and they do not have a large volume and are dense. Softwood pulp from south-
ern regions provides fluffier qualities suitable for hygiene articles such as kitchen towels
or diapers. On the other hand, hardwood pulp has short fibers that increase the paper’s
stiffness, smoothness, and opacity. Since this pulp does not form as many quasicrys-
talline dense areas, a paper’s bulk and absorption capacity are increased.
Lignin is the most reactive substance, followed by hemicelluloses and finally by cel-
lulose. This fact is considered for pulp production [174], [177, Chapter 3]. In practice, two
processes in particular have become established: Kraft cooking with NaOH and Na2 S
(so-called white liquor) and sulfite cooking with H2 SO3 .
From a chemical point of view, bond fractures aim to create lignin fragments in
chemical pulp production. Then, charged functional groups are added to keep them in
solution to separate the cellulose fibers. However, since the degradation processes are
not 100 % lignin specific, the desired carbohydrates are likewise gradually attacked and
degraded. Therefore, the process must stop when there is still some lignin left to avoid
excessive loss of carbohydrates, especially cellulose. At this point, unbleached Kraft pulp
(UKP) is present with the typical brown color of packing paper. After cooking, oxida-
tive delignification with oxygen again significantly reduces the lignin content remaining
from cooking before final bleaching, resulting in bleached Kraft pulp (BKP), ▶Figure 6.23.

Kraft cooking
Kraft cooking is the dominant process for producing chemical pulp [174, Chapter 5].
Reagents of the process are hydrogen sulfide anions HS⊖ for fragmenting the lignin and
hydroxide anions HO⊖ for dissolving the fragments.
In the nineteenth century, Kraft cooking served only to produce very thick un-
bleached papers, and sulfite cooking was the preferred process. The reason was the
dark color of Kraft pulp. During the Second World War, however, a process was devel-
oped to efficiently bleach Kraft pulp with chlorine dioxide. Kraft cooking became the
predominant process with this method and its superior technological advantages.
The reaction runs in three phases. In the first phase, lignin and carbohydrates with
comparable reactivity are degraded and dissolved. In the second phase, about 90 % of
lignin is selectively extracted. The remaining lignin portion can only be further removed
in the third phase at the expense of large parts of the carbohydrates (hemicelluloses)
since carbohydrates exhibit high reactivity again. Kraft cooking must therefore stop
when the last phase is reached. ▶Table 6.4 displays typical proportions before and after
Kraft cooking. The result is light brown pulp, generically referred to as unbleached Kraft
584 � 6 Structure of paint systems

Table 6.4: Exemplary amounts of cellulose, hemicelluloses, and lignin in untreated wood and after Kraft
cooking [174, Chapter 6.5].

Fiber Pine Birch


Wood/% Kraft pulp/% Wood/% Kraft pulp/%

Cellulose 41 38 40 34
Glucomannan 17 4 3 1
Xylan 8 5 30 16
Other carbohydrates 5 – 4 –
Lignin 27 3 20 2

pulp (UKP) or, more specific, unbleached Kraft softwood pulp (UKSP) or unbleached Kraft
hardwood pulp (UKHP), ▶Figure 6.23.
For extraction, the lignin network must be fragmented and provided with func-
tional groups that hold the fragments in the solution until the pulp is separated. Frag-
mentation occurs with HS⊖ in an alkaline solution on the main bonds of the individual
phenylpropane units, the so-called β-O-4 structure. ▶Figure 6.24 represents the primary
reaction.

Degradation of carbohydrates via the peeling reaction


The carbohydrates present in the wood suffer an approximate 10 % loss of substance
during Kraft cooking due to the peeling reaction and the alkaline hydrolysis [174, Chap-
ter 5.6].
The peeling reaction occurs on polysaccharides bearing substituents at the 4-posi-
tion, such as glucomannans, xylans, and cellulose. From the reducing end, a C6 unit is
cleaved from the cellulose through ring-opening and β-elimination. The remaining cel-
lulose chain, shortened by one unit, carries a new reducing end and can again undergo
the peeling reaction. The cleaved C6 unit rearranges to the stable isosaccharinic acid,
which remains in the solution:

The result of the reaction is depolymerization. At a specific point in the degradation, the
residual polymer goes into solution. In particular, glucomannans can be depolymerized
entirely and dissolved with a polymerization degree of about 100 units. Because of its
length, cellulose with a degree of polymerization more than ten times higher, despite
6.7 Paper � 585

Figure 6.24: Main reaction in the Kraft process for dissolving lignin: cleavage of the β-O-4 structure by
hydrogen sulfide HS⊖ . In the alkaline environment, equilibrium between the β-O-4 structure I and the
quinone methide II is favored. In the presence of HS⊖ , a further equilibrium is established between the
quinone methide II and a thioalcohol III, which corresponds to the educt I, but is more nucleophilic. In a
nucleophilic substitution of a phenol unit by the thioalcohol, a thiirane IV and a phenol form irreversibly.
Thus, the lignin network breaks at this point. Elimination of sulfur from the unstable thiirane leads to prod-
ucts such as coniferyl alcohol [174, Chapter 5].

depolymerization, shows no significant dissolution loss due to the peeling reaction. On


xylans, the peeling reaction is limited by the substituents.

Degradation of carbohydrates via alkaline hydrolysis


At the maximum process temperature, alkaline hydrolysis of cellulose increasingly ap-
pears as a side reaction [174, Chapter 5.6]. By unspecific attack of a hydroxide anion,
cellulose is strongly fragmented. ▶Figure 6.25 shows an exemplary reaction course, but
the hydrolysis proceeds in a heterogeneous manner.
586 � 6 Structure of paint systems

Figure 6.25: Cleavage of a cellulose chain by alkaline hydrolysis [174, ch. 5.6]. The chain is fragmented
nonspecifically, giving rise to a reducing end. The reaction can be run repeatedly, fragmenting the cellulose
significantly.

Oxidative delignification and final bleaching of Kraft pulps


The small amount (2–5 %) of residual lignin after Kraft cooking gives the pulp its typi-
cal brown color, as we know from packing paper [174, Chapters 9, 10]. To produce high-
quality white paper, exhaustive delignification and final bleaching to the desired bright-
ness must now occur. A series of oxidations and extractions achieve this until an almost
lignin-free pulp is obtained. In principle, the process is based on the oxidative ring-
opening of phenols and their conversion into soluble carboxylates:

An efficient oxidizing and bleaching agent was chlorine, which disproportionates in wa-
ter to form hypochlorite ClO⊖ . After a process for producing chlorine dioxide ClO2 was
found around 1940, the process sequence CEDED (chlorine, alkaline extraction, chlorine
6.7 Paper � 587

dioxide) could lead to a fully bleached sulfate pulp. The CE-stage is the initial, actual ox-
idation stage, which performs the delignification, while the D-stage achieves the final
whitening.
Around 1970, oxygen slowly replaced chlorine as an oxidizing agent (O-stage) be-
cause of the discovery that small amounts of magnesium salts could protect the pulp
from unwanted oxidative degradation. An O-stage can remove about 50 % of lignin. At
the same time, chlorine dioxide gradually replaced chlorine. This development was ac-
celerated by discovering the highly toxic dibenzodioxins, dibenzofurans, and several
lower chlorinated aliphatics and aromatics in paper mill effluents. By about 1990, there-
fore, all C-stages were rapidly replaced by bleaching stages based on oxygen (O-stage),
chlorine dioxide (D-stage), or hydrogen peroxide (P-stage). Since the desired target pa-
rameters, low residual lignin content, and high brightness cannot be achieved in a sin-
gle step, several stages are combined to form process chains or bleaching sequences in
practice. Typical sequences for the nowadays strongly demanded chlorine-free bleached
pulps elemental chlorine-free ECF and total chlorine-free TCF are

ECF: OD (OP)DD, TCF: O Q(OP)Q(PO)

The initial OD or O-stages support oxidative delignification, in which most residual


lignin is destroyed, followed by the final bleaching or whitening toward the target value.
Typical process parameters are for the
– O-stage: O2 gas, NaOH aq., Mg salts, 30–60 min at 90–100 ℃ and pH 10–11
– P-stage: H2 O2 , 1–3 h at 80–110 ℃ and pH ≈ 11 (transition metals must first be re-
moved in a Q-stage by complexation)
– Q-stage: EDTA or DTPA, 5 min–2 h at 50–90 ℃ and pH 4–7
– D-stage at the beginning of a bleaching sequence (D- or OD-stage): ClO2 , 30–45 min
at 65–75 ℃ and pH 2–3 (ClO2 efficiently degrading lignin; no bleaching is achieved
in the process)
– D-stage at the end of the bleaching sequence: ClO2 , 3 h at 70–80 ℃ and pH 3.4–4.5
(ClO2 efficiently lightening the lignin-poor pulp)

The result is light-colored pulp, which is generally called bleached kraft pulp (BKP) and
marketed as bleached softwood kraft (BSK) or bleached hardwood kraft (BHK). Depend-
ing on the species and origin of the wood, further sales types are derived from these
basic types, ▶Figure 6.23. If bleaching is carried out without elemental chlorine or com-
pletely chlorine-free, the pulp is additionally classified as ECF or TCF pulp.

Sulfite process
Digestion of the wood material is carried out in the sulfite process with CaSO3 , MgSO3 ,
Na2 SO3 or (NH4 )2 SO3 under acidic conditions [174, Chapter 5.11]. The central reaction is
the sulfonation of a phenylpropane unit with SO2⊖ 3 , increasing the solubility with each
588 � 6 Structure of paint systems

additional sulfonic acid function. In the strongly acidic pH range, a high degree of sul-
fonation can be achieved; possibly, acidic hydrolysis of ether bonds in lignin and car-
bohydrates occurs as a side reaction. The result is raw pulp, marketed as unbleached
softwood sulphite (USS) for softwood, ▶Figure 6.23.
Specifically, sulfonation starts with an initial protonation of a benzyl alcohol
group II, followed by elimination of alcohol and addition of the sulfite anion, producing
a sulfonic acid III, ▶Figure 6.26(a).

Figure 6.26: Lignin-dissolving reactions in the sulfite process [174, Chapter 5.11]. In the acidic pH range,
the reaction sequence starts with the protonation of a benzyl alcohol group, followed by the elimination of
alcohol, and finally, the addition of a hydrogen sulfite anion HSO⊖3 .
6.7 Paper � 589

In the alkaline pH range, the phenolic hydroxyl groups of I are involved in the reaction,
▶Figure 6.26(b). The alkaline catalyzed elimination of an alcohol leads to the quinone
methide II, which adds a sulfite anion and becomes the sulfonic acid III. Through the
attack of a further sulfite anion, the lignin network of III is broken, and a disulfonic acid
IV and a new phenol end group form.
Acidic hydrolysis of the carbohydrates also occurs during sulfite cooking, especially
in the case of glucomannan and xylan losses of up to 70 % are possible.

Oxidative delignification and final bleaching of sulfite pulps


Residual lignin also remains in the pulp obtained during sulfite cooking. However, the
pulp is lighter in color than raw Kraft pulp and more accessible to bleach. For TCF grades,
e. g., sequences are used such as

(EOP)Q(PO)
ZEP
(EO)P

(Z-stage: ozone bleaching, E-stage: alkaline extraction). The result is light-colored pulp
sold as bleached softwood sulphite (BSS) or bleached hardwood sulphite (BHS), ▶Fig-
ure 6.23. If the bleaching is done without elemental chlorine or is completely chlorine-
free, the pulp is additionally classified as type ECF or TCF.

6.7.3 Composition and manufacture of paper

As a starting product of paper production serves a pulp of the fiber material mentioned
before [175, 176], [177, Chapter 6.1], [822]. Depending on the desired paper grade, the
mash contains rags or linters, mechanical pulp, or cellulose. The paper pulp shows a
more or less sizeable anionic charge, which is essential for further processing and the
fixation of substances. The causes of the charge are [178, Chapter 3.3]:
– the small natural proportion of carboxyl groups in cellulose
– the natural proportion of acidic hemicelluloses, which accounts for a large part of
the activity
– the high proportion, depending on processing, of phenolic hydroxyl and sulfonic
acid groups in the lignin content
– a variable proportion of degradation products of cellulose formed during bleaching

The raw material mash, which has a solids content of about 0.1–1.5 %, is poured through
a gap onto sieve-like screens that rapidly move under the discharge gap. In the screens,
most of the water flows off, the solids content increases to approximately 20 %, and the
fibers become matted. This process is supported by subsequent rolling and pressing be-
tween several rollers, removing the residual moisture. During pressing, felts or stencil
590 � 6 Structure of paint systems

grids can imprint a texture, e. g., a rough, matte, linen-like, or ribbed one. In producing
high-quality artists’ papers, round screens are also used. The result is raw, highly ab-
sorbent paper with an uneven surface that, in general, will then be sized, coated, and
possibly calendered.
Depending on the intended use of the paper, additives in the paper pulp adapt the
properties of the paper. These include
– fillers
– binders
– retention aids
– dry-strength additives for improvement of stability in the dry state, wet-strength ad-
ditives for the same in the wet state
– dispersants to support a homogeneous pulp
– pigments, dyes, and fixatives, if bulk coloring is desired
– glue, if bulk sizing is desired

It is possible to add the necessary chemicals directly to the pulp (bulk sizing and col-
oring) rather than in a subsequent separate stage (surface sizing and coloring). Many
additives discussed below have more than one function in the papermaking process;
▶Table 6.5 contains an overview.

Fillers
Fillers are finely ground mineral and synthetic substances added to the paper pulp. They
change the properties of the paper to a large extent [177, Chapter 6.1], [179, Chapter 2.2.1].
An important use case is to increase the quantity of paper by adding cheap fillers to
reduce the product’s price. Increasingly important is the improvement of properties
such as
– optics of a paper sheet (brightness/whiteness, opacity/transparency in wet and dry
state, tinting, or coloring)
– structural condition (dimensional stability, weight)
– surface texture (smoothness, uniformity, printability, writability)

Suitable materials are kaolin Al4 (OH)8 (Si4 O10 ), natural calcium carbonate (chalk, cal-
cite, limestone, ground calcium carbonate GCC), or precipitated calcium carbonate
(PCC), talcum Mg3 (H2 O)(Si4 O10 ), silicon dioxide in various forms, calcined alumina
Al2 O3 /SiO2 , and other aluminum-magnesium silicates, barite or synthetic barium sul-
fate, gypsum, anhydrite, alumina Al(OH)3 and Al2 O3 ⋅ H2 O.
In addition to light-colored fillers, white pigments such as titanium dioxide are often
added to the white paper to increase its brightness and opacity further. This admixture
is particularly necessary for papers that must be opaque white, even in wet or oily envi-
ronments. Furthermore, with white pigments, high scattering is desirable. For pigments
and fillers alike, the following applies: the higher the refractive index to air, water, and
Table 6.5: Functional ranges of various polymer additives in the context of papermaking. +++: frequent use, +: also used, Mr : molecular weight of the polymer, Q: charge
of the polymer.

Polymer Binder Strength additive Retention aid Fixative Sizing agent Dispersant

Starch
– cationic +++ +++
– anionic +
– neutral +++ + +
Cellulose
– CMC +++ + +++ +
Gums +++ +++
PAM anionic +++ (Mr medium, Q medium) +++ (Mr high, Q low-medium) +++ (Mr low, Q high)
PAM cationic +++ (Mr medium, Q medium) +++ (Mr high, Q low-medium) +
PAAE resin +++ +
PVOH-PVOAc +++ + +++ (PVOAc high)
PVAm-PVF +++ +++ (Mr high) +++ (Mr low, Q high)
PEI +++ +++ (Mr high, Q high) +++ (Mr low, Q high)
pDADMAC +++ +++ (Mr high) +++ (Mr low, Q high)
PEO +

6.7 Paper
� 591
592 � 6 Structure of paint systems

binder, the more opaque the pigment, ▶Section 1.6.8 at p. 67. Since fillers such as chalk or
kaolin, as well as the pulp, have a refractive index of n ≈ 1.6, they are opaque white only
in the air (n = 0) but not in aqueous, oily, or resinous media, whose refractive indices
are also around 1.6. For wet opaque papers, therefore, highly refractive, although more
expensive, white pigments such as anatase (n = 2.5) or rutile (n = 2.7) are necessary.

Colorants
Colorants (color-strong pigments or dyes) can be added to the paper pulp to color the
whole mass. Organic pigments are transparent; inorganic pigments have an opaque ef-
fect due to their scattering capacity. They offer insolubility, lack of migration tendency,
and fastness to solvents, which originate from printing inks. Typical inorganic pigments
are iron oxides (yellow, orange, red, brown, black), chromium(III) oxide (green), soluble
iron blue (blue, in the paper pulp), insoluble iron blue (blue, as a coating), and mixed ox-
ide pigments (many colors), in addition to organic azo and phthalocyanine compounds,
and carbon black [203, keyword “paper”]. Current assortments of colorant suppliers con-
tain mainly organic azo and phthalocyanine compounds. The binding of the colorants
to the fiber takes place in several ways:
– mechanical retention through the paper felt
– adsorption on the fiber by van der Waals forces and ionic bonding
– entrapment by other particles (colloids such as aluminum hydroxide or resin pre-
cipitates)
– flocculation, i. e., by binding to the fiber using a retention aid

The dyeing of paper and the dyes suitable for this purpose are discussed in detail in
▶Section 5.2. Adhesion mechanisms between the dye and the paper fiber are the subject
of the following sections, separated by dye class.
Brighteners raise the whiteness of the paper by fluorescence. Alternatively, ultra-
marine blue compensates for a yellowish tint of the paper pulp.

Retention aids
At the production beginning, the pulp is a very dilute, aqueous mixture with about
0.1–1.5 ̇% solids (cellulose, fillers). Then it gradually is concentrated by sieving and press-
ing until it becomes a moist paper felt. In the process, most of the constituents can be lost
with the water through the comparatively coarse-meshed screens (0.1-millimeter range).
Since some fillers are expensive, manufacturers avoid removing them from the pa-
per pulp during filtration with the wastewater by adding retention aids, [822], [177, Chap-
ter 6.7], [826, Chapter 14], [180, p. 76ff], [181, Chapter 17]. Together with the anionic cellu-
lose on the one hand and the fillers, on the other hand, these agents form flocs. They are
easier to settle or filter in the wires, and thus remain in the paper to a greater extent.
Initially, the partners are only loosely held together by ionic bonds. However, polymeric
bridging agents attach themselves more tightly to the partners and draw the particles
together until they also come into contact and form bonds. If the paper pulp contains
6.7 Paper � 593

Figure 6.27: Polymers used as retention aids in papermaking [822], [177, Chapter 6.7], [826, Chapter 14],
[180, p. 76ff], [181, Chapter 17].

fines, these initially form small mobile “sticky” balls, which bind the larger particles
together. Chemically, retention aids can have different compositions:
– They consist of silica or silicate-based microparticles made from water glass.
– They can be built from aluminum salts and inorganic compounds such as aluminum
sulfate, sodium aluminate, poly aluminum chloride (PAC), and bentonite. For the
production of acid-free paper today, chalk (CaCO3 ) is preferred.
– Polymers are applicable, forming long bridges between partners and becoming
chain-like with higher molecular weight; ▶Figure 6.27 shows examples. Some
594 � 6 Structure of paint systems

polymeric retention aids such as polyDADMAC or PEI can also act as fixatives or
strengthening additives; see below. However, unlike these, they have a high degree
of polymerization, and thus can form the necessary bridges between the partners.
In addition, for fixation on the fiber, they are usually weakly cationic:
– cationic starch, cationic galactomannans (guarans), carboxymethyl cellulose
(CMC)
– polymeric diallyl dimethyl ammonium chloride (polyDADMAC)
– polyacrylamides (PAM) with high molar mass (10 million), used prevalently
cationic modified, also neutral or anionic
– polyethylenimine (PEI), neutral or cationic
– polyamideamine (PAmA), polyethylene oxide (PEO)
– polyvinylamine (PVAm), due to production by incomplete hydrolysis also as a
copolymer with polyvinylformamide PVAm-PVF

Anionic polymers such as anionically modified polyacrylamide are used less frequently.
However, they are effective because they can form strong hydrogen bonds to cellulose.
The aluminum salts frequently used as flocculants until the last century, mostly alu-
minum sulfates, hydrolyze slowly in the pulp and even later. In the process, they form
sulfuric acid, so the paper becomes acidic. This acidification is of great concern from
an archival point of view since acids promote paper decay, ▶Section 6.7.8. Therefore,
calcium salts such as chalk (CaCO3 ) are employed today. They remain in the paper as an
alkali reserve to slow future acidification when used in excess.
polyDADMAC is a polymer of diallyl dimethyl ammonium chloride (DADMAC) and
possesses a permanent cationic charge through which it adsorbs to anionic fibers and
particles:

Polyacrylamides are neutral by nature but are usually cationic modified by copoly-
merization with acryloxyethyl trimethyl ammonium chloride or trimethyl ammonium
propylacrylamide. Anionic PAM is obtained by copolymerization with acrylic acid. De-
spite its charge, which corresponds to cellulosic fibers, the adhesion is good since the
amide form strong hydrogen bonds with cellulose.
Polyethylenimine represents a polymeric secondary amine with a high number of
charged sites. Therefore, the effective charge of this agent is higher, the lower the pH
value is
6.7 Paper � 595

Strengthening additives
The composition of paper significantly determines its strength. If we have a choice,
we can use longer cellulose fibers, add fewer fillers, and avoid acidic pH values to in-
crease the strength properties of the paper. If this is not possible, additives will help:
Dry-strength additives increase dry strength, tear strength, and abrasion fastness in dry
paper and allow the use of lower-quality pulps [176, Chapter 6], [178, Chapter 5], [177,
Chapter 6.3], [179, Chapter 3.6.5], [826, Chapter 16], [181, Chapter 13].
Wet-strength additives increase the strength in the wet state [176, Chapter 7]. The
additives are necessary for combinations of properties that would otherwise be diffi-
cult to obtain, such as high tear strength at high porosity, where bonding would not be
an option, or tear resistance in the wet state. A wide range of substances is applicable,
▶Figure 6.28:
– cationic modified starch
– unmodified starch, starch paste, modified starch (oxidatively or enzymatically de-
graded), anionic starch (*)
– vegetable gums
– anionic (*) and cationic polyacrylamides (PAM)
– soluble cellulose derivatives such as CMC (*)
– synthetic polymers such as polyethyleneimines (PEI), polyvinylamine-polyvinyl-
formamide copolymers (PVAm-PVF), or polyvinylalcohol-polyvinylacetate copoly-
mers (PVOH-PVOAc)

The starred agents (*) require the addition of cationic fixatives. The purpose of the ad-
ditives is:
– to maintain existing bonds between the cellulose chains and fibers (e. g., hydrogen
bonds) by preventing the fiber from swelling when wet and losing the close contact
between the cellulose strands
– to form further (water-resistant) bonds, ideally of a covalent nature

The additives diffuse into the fibers and between the cellulose chains to protect existing
bonds. There, they intertwine and cross-link with themselves and the carbohydrates,
fixing the chains, even in the presence of water. In addition, the additives react with
hydroxyl or carboxyl groups of the carbohydrates to form new, covalent bonds.
All agents have in common a polymeric character with long water-soluble and
preferably cationic chains to build bridges between cellulose fibers and cross-link
596 � 6 Structure of paint systems

Figure 6.28: Polymers used as strengthening agents in papermaking [178, Chapter 5], [177, Chapter 6.3],
[179, Chapter 3.6.5], [826, Chapter 16], [181, Chapter 13]. The structure of starch is shown only schemati-
cally.
6.7 Paper � 597

them. Bonding to cellulose occurs through van der Waals forces, hydrogen bonds, and
ionic bonds.
Starch is one of the oldest agents for strengthening paper and is still widely used
today. However, unmodified starch cannot be applied since native starch does not show
any adhesive effect and does not adsorb to the fiber material. The reason is the structure
of the starch molecules. In native starch, they form compact grains insoluble in water, re-
sulting in a thin suspension with no adhesive effect since the supply of starch molecules
cannot be set into effect. Starch is, therefore, first opened up. In purely physical terms,
hot steam achieves this opening up, which causes the starch granules to swell and burst
due to water absorption so that the released starch molecules gelatinize. In the process,
part of the starch, amylose, dissolves, and the viscosity of the suspension considerably
increases while crystalline parts of the starch structure melt. In essence, hydrogen bonds
between glucose molecules are cleaved, ▶Figure 6.17 [280, Chapter [Link]]. The exact
process occurs in the thickening of sauces: Starchy thickeners must boil before they de-
velop their adhesive effect.
The following applications result (▶Table 6.6):
– use of the (neutral) starch paste
– production of cationically modified starch
– production of anionically modified starch

Neutral or anionically modified starch bound, for the most part, only reversible to the
cellulose fiber. Cationically modified starch possesses the most significant strengthening
effect; approximately 1–5 % of the primary hydroxyl groups are cationized. It is irre-
versibly adsorbed on the cellulose fiber; besides hydrogen bonds, mainly ionic bonds
between starch cations and the anion activity of the cellulose are involved in the bind-
ing.
The effectiveness of starch resides in the fact that it provides new binding sites af-
ter its binding. In addition, it effectively increases the number of fiber-fiber hydrogen
bonds, which require sizeable spatial proximity of the partners (typical distances are
0.3 nm). This proximity is rarely achieved in the coarse fiber felt, but starch can fill the
wide gaps between the fibers with a bonding matrix and cross-link them.
Quaternary ammonium salts or protonated tertiary amines act as cations, which are
introduced into the starch molecule by reactive monomers such as N,N-diethylamino-
ethyl chloride, or 2,3-epoxypropyl trimethylammonium chloride:
Table 6.6: Starch and key technical modifications [277, 278, 280].

Modification Charge Molar mass Viscosity Reagent Example

Native None High High


– gelatinized None High High
Cationic Cationic High Lower 2,3-Epoxypropyl trimethylammonium
chloride, N,N-diethylaminoethyl chloride

Anionic Anionic High Lower NaH� PO�


598 � 6 Structure of paint systems

Etherified None High Lower Ethylene oxide, propylene oxide

Oxidized Anionic High, possibly medium Lower NaOCl, H� O� , MnO⊖�

Degraded (hydrolyzed)
– thin boiling None High–medium Lower Mild acid HCl diluted
– thermal None Medium–low Lower H� PO� , H� SO� , HCl dilute
– pyrolytic None Medium–low Lower Heat, dry, some acid Glucose, maltose, higher saccarides
– enzymatic None Medium–low Lower Enzymes
6.7 Paper � 599

The direct use of the high molecular weight neutral starch paste encounters process-
ing problems due to its high viscosity. Consequently, further processing steps are re-
quired, leading to products with lower viscosity and better suitability for machine ap-
plication [277, p. 50ff, Chapter 6.1], [278]. The treatment with enzymes, heat, or dilute
acids hydrolytically degrades starch to medium and small fragments (dextrins). Expo-
sure to sodium phosphate leads to phosphoric acid starch esters of anionic character.
Thermally assisted oxidation with hydrogen peroxide, hypochlorite, ammonium persul-
fate, or permanganate converts the hydroxyl groups at C6 , C2 , and C3 (possibly opening
up the glucopyranose ring) into aldehydes or anionic carboxyl groups, depending on the
pH value. Starch can significantly degrade during the process depending on the reaction
procedure’s settings [277, 829].
The adsorption of neutral or anionic starch onto the fiber material, and thus the
stiffening effect increases if starch couples to the fiber with cationic polymers such as
protonated polyvinylamine [179, Chapter 3.4.2].
Plant-derived gums such as guaran and locust bean gum are predominantly used.
The basic structure of these gums, with a molecular weight of 200 000–300 000, is a chain
of mannan with short side chains of galactose [280, Chapters [Link], [Link]]. They are
employed for unbleached Kraft papers. The hydrophilic gums are similar in structure
to cellulose. Adsorbed through van der Waals forces and numerous hydrogen bonds
to the fibers, they have a strengthening effect by increasing the contact between the
fibers after drying, reinforcing the fiber-fiber bonds, and improving their formation. To
a certain extent, gums also have a cross-linking effect.
Since natural plant gums are nonionic, they are not optimally adsorbed on the fiber.
Therefore, cationic modified derivatives are preferred.
Polyacrylamides (PAM), which we have already met as retention aids (see above),
contain about 10 % ionic monomers in addition to acrylamide. Cationic PAM are copoly-
mers with ammonium salts such as acryloxyethyl trimethyl ammonium chloride or
methosulfate H3 COSO2 O⊖ , as well as DADMAC, 3-acrylamido-3-methylbutyl trimethyl-
ammonium chloride, and vinylbenzyl trimethylammonium chloride. Anionic PAM are
obtained as copolymers with acrylic acid and require a cationic fixative to bind them
to the anionic fibers. Possible fixatives include aluminum sulfate or cationic polymers
such as polyamidoamine-epichlorohydrin resins (PAAE resins, see below), serving as
wet-strength additives.
600 � 6 Structure of paint systems

PAM have a strengthening effect, as they offer new, flexibly positioned binding sites
for fiber-fiber hydrogen bonds when the distance between the fibers is too consider-
able for forming direct hydrogen bonds. Hydrogen bonds between the amide group and
cellulose hydroxyl groups are even more potent than those between cellulose hydroxyl
groups themselves:

The acrylate chain is long enough to provide the necessary binding sites for the desired
strength and bridge the distance between fibers. However, it is not so long that it bridges
fibers widely itself. However, bridging is possible; PAM then acts as a retention aid.
Soluble cellulose derivatives such as CMC also act as fixatives through their long
chains and require a cationic fixative like PAAE resin.
Polyvinylamines (PVAm) form from polyvinylformamide (PVF) by hydrolysis so that
they can be used alone or as copolymer p(VAm-VF). PVAm forms hydrogen bonds with
cellulose via the amino group and has a cationic effect in acidic conditions as an am-
monium salt. PVAm can also be used as a retention aid or a fixative, depending on the
electrical charge and chain length.
Polyethylene imines (PEI) can also be used as retention aids or fixatives, depending
on its electrical charge and chain length.
Polyvinylalcohols (PVOH) form from polyvinylacetate PVOAc by hydrolysis and act
as copolymer p(VOH-VOAc). The higher the remaining proportion of PVOAc, the more
hydrophobic the polymer and the higher its suitability as a sizing agent.
PAAE resins belong to a class of polymeric compounds that are used, among other
things, as wet-strength additives [176, Chapter 7]. These additives consist of a polymeric
backbone that is cationic functionalized. Three types of polymers with amine functions
functionalized with epichlorohydrin are chiefly used:

Poly(amido amine) Poly(alkylene polyamine) Polymer amine Reactive group


▶ PAE/PAAE ▶ PAPAE 3-Hydroxy-azetidinium
chloride
▶ PAE/PAAE ▶ APE Glycidyl-ammonium,
(2,3-epoxypropyl)ammonium

Functionalization of the secondary amino groups of the polymer with epichlorohydrin


leads to equilibrium in an aqueous solution between the open-chain amino chlorohy-
drin form and the cyclic azetidinium chloride form:
6.7 Paper � 601

The azetidinium form has the necessary cation charge and a high ring tension, making
it very reactive. The reaction of tertiary amino groups of the polymer yields cationic
glycidylammonium compounds.
PAE or poly(amidoamine)-epichlorohydrin resins, PAAE are the most important ad-
ditives of this group. They are prepared from epichlorohydrin and a polyamidoamine;
the latter is derived from a dicarboxylic acid and a polyalkylene polyamine. Common
components are adipic acid and ethylenediamine or diethylenetriamine:

PAAE resins are cationic, possessing charged azetidinium groups due to epoxidation.
The resins are reactive due to the high ring tension and can form covalent bonds (▶Fig-
ure 6.29) with
– other polymer chains (self-cross-linking)
– cellulose chains, especially the carboxyl groups (cross-linking)
– water
602 � 6 Structure of paint systems

Figure 6.29: Reaction possibilities of a polymer amine-epichlorohydrin (ECH) resin in the azetidinium
form [176, Chapter 7]. Cross-linking opportunities are particularly interesting to the function as a strength-
ening agent (wet-strength additive).

After the paper is scooped, the cross-linking products contribute to the strength in both
dry and wet states.
PAPAE resins (poly(alkenylpolyamine)-epichlorohydrin) are products of the direct
reaction of polyalkenyl polyamines with epichlorohydrin; APE resins are those of the
reaction of amine polymers with epichlorohydrin.

Fixatives
Fixatives constitute an important part of the pulp [203, keyword “paper”], [177, Chap-
ter 6.8],[179, Chapter 3.6.1], [826, Chapter 14], [180, p. 74ff]. Their task is to permanently
fix ionic fillers and dyes to the fiber by forming ionic bonds. This function is crucial
when, e. g., the dye does not have a natural affinity for the fiber, has too low an affinity,
as is the case with acid dyes, or if the dye charge has incorrect polarity. Such dyes must
be specifically fixed to avoid premature washout during the manufacturing process.
Since the substances to be fixed are generally anionic, cationic fixatives are needed.
They can take effect in two ways. In the first case, the anionic fiber is recharged by cou-
6.7 Paper � 603

pling the fixative to the fiber, giving the fiber a net positive charge on the outside. This
way, a dye anion can couple to this cationic charged fiber. In the second case, the dye
anion is first recharged by forming a cationic complex with the fixative, which then
couples to the anionic fiber:

Fixatives (▶Figure 6.30) are small molecules with a higher charge density than the sub-
stance to be fixed. Therefore, they have low molecular masses and a high number of
charges. Cationic fixatives are
– aluminum salts, especially Al2 (SO4 )3 or poly aluminum chloride PAC, sodium alu-
minate NaAl(OH)4 , formerly alum
– cationic polymers such as polyDADMAC, polyacrylamido epichlorohydrin resins
(PAM resins), and polyamidoamine-epichlorohydrin resins (PAAE resins), retaining
their charge regardless of the pH value
– polyvinylamine (PVAm) and polyethylenimine (PEI), which are charged in the acidic
pH range
– dicyandiamide/formaldehyde resins

We are already familiar with them as retention aids and strengthening agents (see
above).
Some polymeric fixatives such as polyDADMAC, PVAm, or PEI can also act as reten-
tion aids (see above). However, unlike these, fixatives have a low degree of polymeriza-
tion and high charge.
Examples of the less commonly used anionic fixatives commercially offered as poly-
mers are [203, keyword “paper”]:
– methylene-bridged condensation products of arylsulfonic acids and hydroxyaryl-
sulfones (a commercial product is, e. g., MESITOL® )
– condensation products of naphthalenesulfonic acid with formaldehyde, forming
lakes with the dye [827, p. 5] (a commercial product is, e. g., Tamol® )

Condensation products of hydroxyarylsulfones and arylsulfonic acids with formalde-


hyde have a similar structure to the resol resins. Due to the reactivities of the reactants
604 � 6 Structure of paint systems

Figure 6.30: Polymers used as fixatives in papermaking [203, keyword “paper”], [177, Chapter 6.8],
[179, Chapter 3.6.1], [826, Chapter 14], [180, p. 74ff]. Some compounds carry inherent charges, while oth-
ers attain charge at appropriate pH.

(sulfonic acids have a meta-directing effect, phenol groups ortho-/para-directing), we


can expect structures similar to the following ones (the dots marking the reactive sites):
6.7 Paper � 605

Aluminum sulfate, formerly alum, is the best-known agent for fixing resin glues. It hy-
drolyzes in an aqueous solution to form aluminum ions and sulfuric acid:

A pH value of around 3.5 is established during the reaction. These acidic conditions lead
to paper decay (▶Section 6.7.8) and are increasingly avoided today.
Polyethylenimine is a polymeric secondary amine with a high number of charged
sites. The effective charge of this agent is higher the lower the pH value is

Dicyandiamide resins are made from dicyandiamide, formaldehyde, and ammonium


chloride [828, p. 21] and have a permanent cationic charge:

6.7.4 Sizing and coating

The plain paper obtained after pressing and drying has a rough surface and is in-
terspersed with numerous capillary cavities in the fiber felt. In conjunction with hy-
drophilic fibers, these cavities lead to high water absorption and absorbency in plain
paper. Therefore, it can be used as blotting or filtering paper in this form, but not yet
for writing, printing, or drawing.
For such papers, the writing or printing inks or colors must be prevented from
bleeding. Furthermore, to obtain a sharp and clean appearance of the printed or writ-
ten typefaces, they are surface-coated to achieve a defined degree of absorbency, ink
affinity, smoothness, and opacity [177, Chapter 6.2], [188, 822], [181, Chapters 16, 20]. Two
processes are often combined:
606 � 6 Structure of paint systems

Figure 6.31: Influence of sizing on aqueous writing fluids (blue).

– Sizing was already applied to the first papers since it renders the surface hydropho-
bic, reduces capillarity and absorbency, equalizes pore size, and increases the
smoothness of the paper.
▶Figure 6.31 depicts the process. Untreated paper (a) has high absorbency and
quickly absorbs the writing fluid. The fluid runs over a wide area and bleeds out.
Since narrow capillaries strongly attract liquids, differently-sized capillaries in the
paper felt lead to feathering and blurred edges.
The paper is partially hydrophobic in the ideal state, and the ink stands well and
sharply on the paper (b). As a result, the capillary sizes are equalized by the sizing
agent’s partial filling of the pores, and the feathering is reduced.
Excessive hydrophobicity results in repulsion (c): the writing fluid is repelled from
the paper and forms tiny isolated droplets. Such behavior may be desirable for spe-
cialty papers such as oiled or grease-proof paper. Sizing also seals the fiber-fiber
binding sites and increases or stabilizes the strength of the paper.
– Coating is a surface finishing carried out with a particular coating color. The aim
is a closed, smooth surface with uniform pore size, a high degree of whiteness and
opacity, and a certain feel due to fillers and colorants. The coating also increases
writeability or printability.

In both processes, fiber cavities in the paper felt, which act as capillaries, are con-
trolled; large voids are closed with fine-grained fillers or sizing agents, and small ones
are brought to as uniform a size as possible. A minimum absorbency must remain so
that writing or printing ink can penetrate the paper to achieve adequate adhesion. The
hydrophobic coating prevents the excess penetration and bleeding of the predominantly
aqueous or hydrophilic inks. In the case of copying paper, the hydrophobic coating en-
sures good adhesion of the hydrophobic toner. The coating can also be enhanced with
chemical auxiliaries to adhesion to the writing, drawing, or printing inks.
6.7 Paper � 607

Sizing agents
For the hydrophobic modification of paper, we can resort to sizing agents, which often
simultaneously perform the functions of retention aids, fixatives, or strengthening ad-
ditives. Their history is as old as that of paper: Rice starch was already applied as a
sizing agent 2000 years ago. Around 1280, sizing with animal glue and alum developed,
from which the name of the process was derived. It remained the leading process until
1806. Around this time, the innovation of resin soap sizing was introduced; it was not
until 1950 that modern polymers began to appear. Modern agents are presented in ▶Fig-
ure 6.32 [175, Chapter 14], [177, Chapter 6.2], [822], [180, p. 63ff], [179, Chapter 3.6.4], [223,
Chapter 5.2], [181, Chapters 16, 20]:

Figure 6.32: Agents used for paper sizing [175, Chapter 14], [177, Chapter 6.2], [822], [180, p. 63ff],
[179, Chapter 3.6.4], [223, Chapter 5.2], [181, Chapters 16, 20]. Traditional agents such as gelatin or animal
glue are still used for artists’ papers. MS: maleic acid, MSA: maleic anhydride.
608 � 6 Structure of paint systems

– tree resins (bulk sizing)


– reactive monomers such as alkylketene dimer AKD, alkenyl succinic anhydride ASA
(bulk and surface sizing)
– polymers such as styrene-acrylate copolymers, styrene-maleinate copolymers, or
polyurethanes of toluene diisocyanate and glycerol monostearate (surface, but also
bulk sizing)
– starch paste and cationic, oxidized and enzymatically, thermally, or acidically de-
graded starch, ▶Table 6.6 at p. 598 (surface sizing, less used today)

Starch derivatives, however, are used more often as strengthening agents.


The resins are alkaline saponified balsamic resins (▶Section 7.4.8), i. e., alkali salts of
the resin acids abietic acid, levopimaric acid, and dehydroabietic acid. Some resin acids
can participate as a diene in a Diels–Alder reaction with maleic anhydride and react to a
modified resin acid, which contains two additional carboxyl groups for anchoring after
hydrolysis of the anhydride:

The resin acid anions are precipitated by AlIII salts (aluminum sulfate, poly aluminum
chloride, alum) [826, Chapter 17]. Significant to the reaction is a series of complexes that
aluminum forms in an aqueous solution:
6.7 Paper � 609

In the required acidic pH range, resin acid anions build a cationic aluminate complex
ionically adsorbed on the anionic fiber (ABA = abietic acid anion). Zerler and Kästner
[177, Chapter 6.2] and Roberts [178, Chapter 7.3] elaborate this further:

According to [175, Chapter 14.3.2], polynuclear AlIII complexes may also be involved in
binding the cellulose to resin acids, e. g., a trinuclear complex such as structure II:

Polymers can be cationic or anionic modified. For bulk sizing, cationic polymers are
preferred because they adhere ionically to the fiber material. For surface sizing, both
polarities are in use. Polymers act by forming a mosaic of hydrophilic and hydropho-
bic regions on the paper, thus increasing the wettability. When used as copy paper, the
hydrophobic areas can improve the adhesion of the toner resin.
As hydrophobic polyacrylates, copolymers of methyl methacrylate, n-butyl acrylate,
and styrene p(MMA-BA-styrene) are employed. As in acrylic paints, they form a film
and adhere via electrostatic and van der Waals forces to the paper’s surface. Anionic
polymers are obtained with small amounts of acrylic acid or maleic anhydride; cationic
polymers with tertiary or quaternary amino compounds. A well-known example of a
cationic monomer is acryloxyethyl trimethyl ammonium chloride:
610 � 6 Structure of paint systems

Due to the possibility of combining all components in varying amounts, we obtain a


range of polymeric sizing agents:

Anionic: p(AA-MMA-BA-styrene)
p(AA-styrene-maleic acid)
p(AA-styrene-maleic acid monopropylester)
p(styrene-maleic acid-maleic acid monoisopropylester-diisobutylene)
Cationic: p(MMA-BA-styrene-acryloxethyl trimethylammonium chloride)
p(styrene-acrylonitrile-acrylamide)

The transition to styrene-maleic acid copolymers p(styrene-maleic acid) is smooth.


This polymer, possibly with olefins, acrylic acid, and acrylic acid derivatives as further
comonomers, provides an anionic polymer with alkaline pH used for surface sizing.
Polyurethanes made from diisocyanates and hydrophobic diols are also employed,
which can be charged with anionic or cationic monomers. An example is a polyurethane
made from toluene diisocyanate and glycerol monostearate, which can be cationic mod-
ified. In addition, cross-linking with epichlorohydrin yields an anionic product.
The hydrophobic properties of alkylketene dimer (AKD) and alkenyl succinic anhy-
dride (ASA) are due to large hydrocarbon-like sections, which in the case of AKD are
derived from C8−22 -fatty acids and in the case of ASA from alkenes formed by isomeriza-
tion from 1-alkenes:
6.7 Paper � 611

AKD and ASA react with hydroxyl groups of cellulose to form a cellulose ester. Then they
adhere covalently to the paper through the ester bond:

Coating colors
With or without the sizing, a further coating, the coating color, can be applied to the
paper [175, Chapter 16], [203, keyword “paper”, Chapter 9.2], [179, Chapter 3.6.9], [223,
Chapter 5.3], [180, p. 86ff]. Their purpose is to give a smoother surface and a certain look
and feel to the coated paper through fillers. The necessity of the surface smoothness
results from the fact that the dimensions of the paper fibers used are a few millimeters in
length and 50–100 µm in width. In this way, halftone dots from 300 dpi printing already
are the same size as or even smaller than the fibers, so unsmoothed paper limits the
possible resolution severely. White pigments improve the whiteness and brightness of
the paper. Colored pigments are used for coloring; however, about 95 % of paper dyeing
is done directly in the pulp.
It is possible to dispense with coating and add the necessary fillers and pigments
directly to the paper pulp prior to pressing and drying, thus eliminating the need for
a separate step if the chemistry of the auxiliaries permits this. A typical coating color
contains (▶Figure 6.33):
– fillers and pigments as main ingredients
– dispersants for stabilizing the filler, and pigment dispersion
– binders to bind the substances to the paper
– cellulose- or polyacrylate-based thickeners
– other auxiliaries such as wet-strength additives or defoamers

Fillers and pigments are mineral or synthetic substances. Whether a substance is consid-
ered a pigment or a filler depends considerably on its refractive index. Low-refractive
substances serve as fillers. Commonly used are kaolin Al2 O3 ⋅ 2 SiO2 ⋅ 2 H2 O, calcined clay
Al2 O3 /SiO2 , natural and precipitated calcium carbonate CaCO3 (ground calcium carbon-
612 � 6 Structure of paint systems

Figure 6.33: Constituents of a coating color: wetting agents, dispersants, binders, and thickeners
[175, Chapter 16], [203, keyword “paper”, Section 9.2], [179, Chapter 3.6.9], [223, Chapter 5.3], [180, p. 86ff].

ate GCC, precipitated calcium carbonate PCC), talcum MgO⋅4 SiO2 ⋅H2 O, barite or barium
sulfate (blanc fixe) BaSO4 , aluminum hydroxide Al(OH)3 , gypsum CaSO4 ⋅ 2 H2 O, or tita-
nium dioxide TiO2 [179, Chapter 2.2.2].
Wetting agents and dispersants support the grinding of the above mentioned sub-
stances and dispersion in the solvent. ▶Section 6.3 examines the underlying processes
in more detail.
6.7 Paper � 613

Classic agents are sodium or ammonium salts of polyacrylic acid or polyphospho-


ric acid Na5 P3 O10 . Due to their charge, they have an electrostatically stabilizing effect.
Polymers such as polyacrylamide (PAM), starch, or polyvinylalcohol (PVOH) form pro-
tective colloids around the particles and have a sterically stabilizing effect. Both effects
are present in carboxymethyl cellulose CMC and products from lignosulfonic acids and
sulfonic acid-containing aromatic formaldehyde resins, which act electrostatically and
sterically. Some compounds also act as binders, retention aids, or fixatives, see above.
A wide range of natural substances can be applied as binders. They include starch,
casein, soy proteins, and cellulose. Starch and cellulose must be modified to achieve their
full effectiveness; possible candidates are oxidized starch, hydrolyzed starch (including
ester and ether derivatives), and carboxylated and etherified cellulose (CMC, HEC).
Fully synthetic binders have become important. The most significant are styrene
acrylates (SA) and styrene-butadiene latex (SB). Styrene acrylates are copolymers of
butyl acrylate with styrene or vinyl acetate p(BA-styrene-VAc). Styrene-butadiene lat-
ices are used as carboxylated copolymers (XSB); the carboxylic acids are derived from
small amounts of acrylic acid, methacrylic acid, and maleic acid p(styrene-butadiene
AA/MAA/maleic acid). Polyvinylalcohol or polyvinylacetate is also included in binders.
Thickeners adjust the viscosity of the coating color to achieve optimum processabil-
ity. Typical thickeners are polyvinylalcohol (PVOH) and nonassociative thickeners based
on cellulose or polyacrylic acid, especially carboxylated cellulose CMC, polyacrylic acid
copolymers of acrylic acid/methacrylic acid with methyl acrylate, ethyl acrylate, methyl
methacrylate, or acrylamide p(AA/MAA-MA/EA/MMA/acrylamide). From the field of
associative thickeners, representatives of hydrophobically modified polyurethanes
(HEUR), polyacrylic acids (HASE), and celluloses (HMHEC) are employed. We have de-
scribed the mode of action and structures of thickeners in ▶Section 6.4.

6.7.5 Calendering (satinage)

After dewatering with light pressure through felts, the raw paper is naturally rough or
slightly textured through the felts, but in all cases, uneven. Sized and coated paper is
also matte after the treatment with glue or coating.
The term “calendering” describes operations in which the (moist or already dried)
naturally rough matte paper web is passed through various roller systems (calender)
and thereby more or less firmly pressed [175, Chapter 15]. Calendering aims to achieve
a more even surface suitable for writing or printing on it with rigid printing plates or
showing a higher gloss. Objectives are:
– to strengthen the surface to reduce dusting and tearing of fine fibers.
– to reduce the porosity of the paper, lessening the absorbency and run-off of ink. As
a result, the sharpness of a drawing or printed image is increased. Since this re-
duces the local inhomogeneity of the refractive index, the amount of scattered light
produced by the paper and its brightness decreases while transparency increases.
614 � 6 Structure of paint systems

– To decrease surface roughness in the range of millimeters to the size of fiber diam-
eters is essential for printing methods that use rigid printing plates. They require
sufficient contacts of paper and printing plate, such as rotogravure or copperplate
printing. In the case of copperplate printing, artists also like to use soft paper, into
which the printing plate is impressed deeply.
– to reduce surface roughness in the micrometer range to obtain glossy papers.
A smooth surface induces gloss with a homogeneous refractive index, on which
hardly any diffuse reflection or scattering occurs.

Roller systems may consist of two or many rolls, hard or soft, cold or heated. Typical
systems include:
– many hard rollers (up to 17) heated up to 100 ℃ (supercalender) to produce very
smooth paper grades such as WFC, LWC, and SC (roto-engraving quality)
– two soft heated rollers for LWC (offset and newspaper printing)
– four soft heated rollers for SC, LWC (offset quality)
– many soft rollers for smooth SC and LWC types and glossy WFC (all roto-engraving
quality)

Calendering cannot be considered independently but only in combination with sizing


and coating, as we will see in the example of artists’ papers.

6.7.6 Paper grades, general and industrial

Depending on the raw materials, pulp or cellulose types used, and the type and num-
ber of processing steps, a paper manufacturer can produce many different paper types,
which in turn permit different application types. Each step uses its specific conceptual
taxonomy, ▶Figure 6.34.
The classification according to the material basis takes into account the origin of
the pulp and is ordered by increasing paper quality: wood-containing papers—pulp
papers—rag paper:
– Wood-containing papers are largely made from mechanical pulp, i. e., wood pulp or
cut or ground wood, ▶Section [Link]. However, since wood fibers are torn and are
very short in this type of pulp production, the paper has little mechanical strength.
Therefore, for higher-quality papers for magazines and print inserts, particular
types of pulp are added as a counteracting measure. These consist of longer-fibered
cellulose and serve as reinforcement.
Papers containing wood are cheap and often of poor quality. They frequently con-
tain fillers and additives to improve their properties, which we can recognize when
burned, receive ash volumes of about 14 %.
– Pulp papers or wood-free papers consist of pure pulp, i. e., essentially pure cellu-
lose obtained from chemical pulp from which undesirable woody components such
6.7 Paper � 615

Figure 6.34: Four levels in papermaking, from raw material to application. All levels have their specific con-
ceptual taxonomies, often used side by side. Raw materials are applied to produce intermediates standard-
ized for wood and cellulose pulps, and traded worldwide. The industry combines intermediates into pulp
mixtures processed into many grades of paper. The paper grades are also often standardized for multiple
applications.

as lignin have been chemically removed, ▶Section [Link]. Basically, cellulose re-
mains in the structure in which it is already present in the wood, i. e., in the form
of long, well-preserved fibers. Compared to wood-containing paper, the fibers are
long (1–3 mm), stable, and matte well. Since wood-free papers contain no additives,
they yield practically no ash. They are applied, among other things, as high-quality
writing papers, papers for graphic work (commercial printing), filter papers, and
restoration work. Drawing and layout papers are also mostly pulp papers.
– Rag papers were initially made from rags consisting of textile fibers of flax (linen),
cotton, or hemp. The fibers are long, uninjured, and matte very well into durable,
age-resistant papers, so artists often used them as a base for drawing and painting.
Rag papers contain only a few additives. They are still the preferred choice for high-
quality artists’ papers, or, at least specific long-lasting pulp papers are. Since the late
Middle Ages, the availability of rags has been a limiting factor in paper production,
so cotton fibers have been used directly to produce rag paper without going via
textiles. The fiber raw materials yield very pure virgin cellulose fibers, as we can see
when burning. The typical ash volume is only about 2 % of the paper volume [820].

Four main groups exist according to the type of application:


– Graphic papers include all types of printing paper: press products, stationery, office
(copy) applications, and forms. They are available as roll or format paper.
– Packaging papers comprise paper, board, and paperboard, e. g., wrapping paper,
folding boxboard, or corrugated base paper.
616 � 6 Structure of paint systems

– Specialty papers including types of papers and boards for technical and specialty
applications, e. g., wallpaper, filter, or cigarette papers. Of interest to us are the var-
ious types of artists’ papers as well as those for inkjet application.
– Hygiene papers include tissue papers (made of pulp) and crepe papers (made of
wood pulp) for toilet paper, kitchen rolls, and others.

The type of coating also characterizes papers:


– Natural papers are uncoated papers. Their natural porosity makes them less suit-
able for reproducing images but ideal for writing, office, copying, and laser printing
papers. To control the ink flow, they are sized to varying degrees. A light pigmen-
tation below 5–10 g/m2 , e. g., with chalk for smoothing the surface and increasing
brightness, is not considered a coating.
Natural papers come in different varieties:
– Machine-smooth papers that have not undergone surface treatment other than
processing into paper.
– One-side smooth papers are glossed by a smoothing cylinder but not pressed.
– Satinized papers are more glossed by a calender under pressure and ele-
vated temperature. Multiple calendering or a supercalender results in an even
smoother surface.
Artists’ papers are essentially uncoated.
– Coated papers have a smoother and more uniform surface than uncoated papers.
They are therefore ideal for high-quality reproduction of images with brilliant col-
ors. They include:
– Double-sided coated papers, such as LWC (lightweight coated), art paper, and
picture paper.
– One-side coated papers mainly for packaging, such as label paper or folding box
board (type designations are SBB, SUB, FBB, WLC).

▶Figure 6.35 provides an overview of crucial graphical paper grades commercially


available today in a quality-value relation. ▶Table 6.7 contains details of essential pa-
per grades. All art printing and artists’ papers are strictly wood-free and contain rags in
higher grades. The following section will look at artists’ papers in more detail.

6.7.7 Special case artists’ paper

What has been said so far about basic paper materials and properties applies in princi-
ple to commercial paper products such as writing paper, letterpress paper, catalog pa-
per, and most importantly, artists’ papers. However, the latter represents a niche with
no standardized paper grades. The many different grades of paper that are important
for the artist result from the interaction of a few parameters:
– surface sizing
6.7 Paper � 617

Figure 6.35: Modern commercial graphic paper grades. The abscissa symbolizes the quality of the prod-
uct (brightness, surface, printability) and the ordinate the “value” [832]. Left: The numbers indicate the
grammage in g/m2 . Light blue: uncoated paper. White: coated paper. The artists’ papers are shown in
▶Figure 6.36.

– pressing during dewatering, calendering for surface texturing


– material or pulp mixture
– coating (artists’ papers are, in general, natural papers, so they are not coated)
– carefully determined machine parameters (temperature, press pressure, etc.)
– clean water, preferably spring water

We will discuss specialty papers for a particular art technique in connection with a paint
system, but the parameters mentioned already allow some general statements. In ad-
dition to the material, surface sizing, and pressing during dewatering or calendering
are the main factors determining the properties, ▶Figure 6.36 schematically shows the
relationship between these two parameters. ▶Table 6.8 compiles typical properties of
artists’ papers.

Sizing
Artists’ papers are consistently sized in the pulp to ensure strong cohesion of the fibers
and achieve the necessary mechanical resistance. As a rule, AKD is employed as a neutral
size. AKD may react acidically by hydrolysis, but artists’ papers are buffered with 2–4 %
Table 6.7: Overview of important commercial graphic paper grades and their composition or manufacturing characteristics [176, Chapter 9.5], [180, p. 101], [179, p. 98].

618 � 6 Structure of paint systems


Brighter grades contain more chemical pulp. BS: bulk sizing, SS: surface sizing to control ink flow.

Paper grade Grammage [g/m� ] Mechanical pulp Chemical pulp, Kraft Filler Misc.

Newsprint paper SNP, INP 42–49 +++ + - Possibly bleached and pigmented to
reduce transparency
Directory paper DIR 34–45 +++ -
Magazine paper 49–60
– SC +++ + ++ (alumina) Strong satin finish
– LWC ++ ++ ++ (alumina) Light satin finish, coating 8–10 g/m�
– MWC 70–250 ++ ++ ++ 2x coated
Book paper 60–90 +++ + - Worked to bulk by short fibers
Copying, office, letter and 75–160 + +++ +++ (CaCO� ) BS, SS, uncoated, possibly calendered
writing, woodfree, WFU Many short wood fibers (birch), few long wood fibers (pine)
Image printing paper Many short wood fibers (birch), few long wood fibers (pine), 2–3× coated with CaCO� /alumina (standard coating), possibly plastic for
high gloss. Single coated for matte paper, multiple coated and satin for glossy paper
– HWC, HWC-WF 80–250 - +++ +++ (CaCO� ) Art book, high quality brochures
– WFC, WFU 80–250 - +++ +++ (CaCO� )
Art paper - +++ 2–3× coated (high quality)
6.7 Paper � 619

Figure 6.36: Schematic representation of the influence of surface sizing and pressing with dewatering
or calendering on the application properties and areas of use of artists’ papers. All papers are uncoated
and consist of cotton, and lower qualities are made partly or entirely of cellulose. Light blue: application
properties achievable by sizing and surface.

calcium carbonate CaCO3 as an alkali reserve to keep the pH value to 7.5–9.5 and to be
considered resistant to aging according to ISO 9706, ▶Section 6.7.9. Combined with the
alkali reserve, this results in a neutral-alkaline sizing system. In addition, traditional
sizing with animal glue (gelatine) is also typical.
Depending on the intended use of the paper, surface sizing can follow in order to
set essential application parameters:
– A surface sizing increases the surface strength and resistance to mechanical effects
such as:
– the movement of an eraser or a hard pencil (drawing papers)
– the rubbing off of masking rubber or rubbing crepe or the tearing off of adhe-
sive tape (watercolor papers)
– the brush movements when correcting or removing color (watercolor papers)
– A weaker surface sizing slows down, and a stronger one stops the water or ink ab-
sorption of the paper:
– The open time is prolonged, i. e., the paint or ink remains correctable, smudge-
able, or mixable on the paper with other colors (watercolor, layout, and fine-
liner, acrylic painting papers).
Table 6.8: Overview of artists’ paper and its composition or manufacturing characteristics [1004]. Bulk sizing (BS) and surface sizing (SS) control ink flow, increase the
mechanical resistance, and act as a water, solvent, and ink barrier. SS works in conjunction with calendering. Environmentally-friendly ECF or TCF bleached types are used
as high-quality pulp.

Application Grammage [g/m� ] Chemical pulp (pulp) Rag (cotton) Filler Other

General Uncoated, in general BS, SS if applicable, natural gray, matte or satin finish. For chemical pulp types, addition of > �,� % CaCO�
acc. to ISO 9706 to slow down aging, designation “aging resistant”
– simple +++ − CaCO� Study quality
– high quality ++ ++ CaCO�
– highest quality − +++ −
Watercolor 200–640 See above See above See above Rough, matte, satin finish. SS/calendering increases
620 � 6 Structure of paint systems

strength against corrections, rubbing crepe, tape.


Print 130–320 See above See above See above Soft, worked on volume, in general no SS, absorbent to help
solvent whisk away. Surface rough, matte, satin, suitable for
all techniques. Satin especially for woodcut, screen printing;
soft and voluminous especially suitable for gravure and
planographic printing.
Drawing, pastel 80–250 +++ − CaCO� SS increases strength against erasing, glued 2x if necessary.
Fine-grained surface facilitates pigment abrasion of pencil,
charcoal or chalk. High volume facilitates retention of
abraded pigments.
Bristol 180–920 +++ − CaCO� SS, very smooth and white, colors stand well and brilliant.
Clear markings.
Calligraphy, pen & ink drawing 170–250 +++ − CaCO� Strong satin finish, possibly coated. Colors and markings
stand well and brilliantly.
Graphics, illustration 75–220 +++ − CaCO� In general SS, heavily calendered. SS and calendering
increase strength against erasing and form a wet solvent
and ink barrier.
6.7 Paper � 621

– There is a water and color barrier that makes the paper suitable for low viscos-
ity and alcohol-based media (watercolor, acrylic, layout papers) and prevents
the bleed-through of low-viscosity ink (bristol, layout papers).
– A surface sizing reduces or prevents ink soaking and bleeding:
– Colors stand brilliantly on the paper (watercolor, layout, and bristol paper).
– A sharp drawing emerges (pen drawing, layout, bristol paper).
– Conversely, the lack of surface sizing allows solvents or water to absorb quickly into
the paper so that prints dry rapidly (rotogravure, lithographic papers) and large
amounts of water are absorbed and released in a controlled manner (watercolor
paper).

Surface texture
The surface of the artists’ papers is defined when the water is pressed off the pulp using
various felts for pressing, producing rough (more structured) or matte (less structured)
surfaces. A fine-grained or even smooth surface can be achieved by, i. e., hot pressing.
The effects complement those of sizing:
– A smooth surface increases the surface strength and the resistance against mechan-
ical effects; it slows down the paper’s water or ink absorption and reduces or pre-
vents ink from absorbing and running (see above).
– Conversely, a matte or rough surface allows solvents or water to be quickly absorbed
into the paper so that prints dry quickly (rotogravure, lithographic papers) and large
amounts of water are absorbed and released evenly in a controlled manner (water-
color paper).
– A matte or rough surface produces a higher abrasion of powdery or solid paint sys-
tems like pencils or pastels. It can better fix the abraded pigment particles in the
paper’s pores.

Materials
Today, the highest-quality artists’ papers consist of 100 % cotton; cheaper grades com-
prise mixtures of cotton and chemical pulp or up to 100 % chemical pulp paper for study
papers. Some grades are also made from annual cellulose-producing plants such as bam-
boo, blended with small amounts of cotton to make a mixed media paper.
In the Middle Ages, cotton came from rags, i. e., textile rags, but was soon obtained
directly from the cotton plant due to the scarcity of this resource. It forms long outer
seed hairs and short inner seed hairs, ▶p. 575. Artists’ papers are made from both hair
types; however, some manufacturers prefer the inner, short, soft, and clean seed hairs,
called linters. During papermaking, the long fibers of textile cotton can become knotted
and form webs, disturbing the uniformity of the paper. The knotting and felting process
could also occur during later painting and rubbing with a brush, resulting in undesirable
“paper sausages.” The paint application would then make the disturbances in the paper
visible. Linters fibers produce an open volume paper that is soft but robust and is a
moisture buffer for aqueous solvents (water, alcohol-based fineliners).
622 � 6 Structure of paint systems

In pulp-containing papers, in general, pulp types from softwood, which are soft and
pliable, are mixed with hardwood pulps, which are strong and durable, to obtain opti-
mum paper properties.
A critical material factor is the achievable volume of the paper. In a pulp with long
fibers, the fibers can be well parallel and form dense quasicrystalline areas with high
strength but keep the paper volume low. In pulp with a high proportion of short fibers,
fibers can lay crosswise and form an open-pored paper with a 50–100 % higher volume
than a solid long-fiber paper for the same basis weight. High volume implies pleasant
hand feel, stiffness, and a general sense of value. For some printing techniques, the vol-
ume of the paper is also essential from a technical point of view to absorb the binder and
solvent of the printing ink. For example, the absorption and release of water through
the paper are paramount when working with water-based inks. Softwood pulps from
conifers such as spruce are typical sources for long fibers; short fibers come from hard-
wood pulps (birch, eucalyptus) or linters.

6.7.8 Paper decay

Paper is easy to make, but under certain circumstances, it decays just as readily, as
shown by the works threatened with decay today [823–825]. Especially the groundwood
pulp portions of cheap papers rapidly decay to low molecular weight products. Further-
more, even higher-grade papers sooner or later show signs of decomposition since cel-
lulose as a polyether is slowly hydrolytically cleaved by acids, ▶Figure 6.37(a). Cellulose
is also attacked oxidatively, with the primary hydroxyl group at C6 oxidized to the car-
boxylic acid. The oxidation increases the paper’s acidity even without adding acid. It
stabilizes the intermediate of the protonated ether leading to the cleavage of the ether
bond, ▶Figure 6.37(b).
Protons can come from various sources, one of which is acidic inks. Historic iron gall
inks are a classic problem, but modern dye inks can also react acidic. Papers sized with
tree resins and alum (i. e., the majority of papers from the last two centuries) contain
sulfuric acid from the hydrolysis of excess alum:

Another source of decay appears in graphic works, in which green copper pigments
were used. As [421] delineates, copper ions efficiently catalyze the depolymerization of
cellulose. However, the exact course of the reaction is not yet fully clarified; in the case
of gold plating or metal plating with copper-containing metal alloys, various oxidation
states of copper come into action, via which atmospheric oxygen is employed for cellu-
lose oxidation.
6.7 Paper � 623

Figure 6.37: Basic reactions in paper decay [823].

Ancient papers made before around 1850 show comparatively less decay because rags
were used in their manufacture. As already mentioned, these provide longer fibers,
which, when matted, can form crystalline regions of densely packed and ordered cel-
lulose chains, thus significantly increasing the mechanical strength. In contrast, wood-
containing papers of later times are little interwoven due to their short, broken fibers
and disintegrate at only a few breaks in the cellulose chains. On top of that, the chains
are coated with nonpolar lignin, which strongly impedes the formation of crystalline
regions.
624 � 6 Structure of paint systems

6.7.9 Aging-resistant paper

Especially for cultural assets such as paintings, prints, and drawings on paper as sup-
port, it is of utmost importance that the support lasts as long as possible and does not
fall victim to the described decay phenomena or animal (insects) and plant (mold) pests.
While this was supported in earlier times by the choice of rag as the paper base, today,
we have to pay close attention to how they are composed and what resistance the man-
ufacturer gives them.
In Germany, e. g., two sets of standards exist parallel and with slightly different ob-
jectives. The national standard DIN 6738 “Paper and board, Service life classes” defines
four quality or life classes (LDK). Samples are tested with simulated aging tests: exposing
papers to a temperature of 80 ℃ and relative humidity of 65 % for up to 24 days. Then the
difference between mechanical paper properties before and after aging is determined.
In order to be classified in the highest class LDK 24-85, the papers must still have at least
85 % of the property values, based on the difference. They may then be called aging re-
sistant.
The international standard ISO 9706 “Information and documentation—Paper for
documents—Requirements for permanence,” on the other hand, defines quality char-
acteristics that a paper must meet in its original state. These include specific mechani-
cal properties, a certain alkali reserve (usually as a calcium carbonate buffer CaCO3 ), a
high oxidation resistance, and a slightly alkaline pH of 7.5–10. Oxidation resistance, ex-
pressed as a kappa number, is tolerated to a maximum lignin content above this number
but below 1 %.
Artists’ papers are in general certified by manufacturers in accordance with ISO
9706 as aging resistant to museum standards. In a direct comparison of the two sets of
standards, official state archives and archive administrations regard only ISO 9706 as
suitable for the permanent protection of cultural property [1006]. It is stated that the
ISO 9706 sets stricter standards for the paper than DIN 6738, which itself also speaks of
life span estimates, and on top of this, refers in a note to the criteria of ISO 9706. Thus,
the lignin content according to DIN may be much higher than according to the ISO and
the proportion of acid waste paper.
To summarize, several practical and regulatory approaches exist for assessing pa-
per resistance against aging, so the suitability of paper, treatment, and storage must be
evaluated per case. Regardless of the certifications of the paper, storage conditions sig-
nificantly determine its actual preservation.

6.7.10 Yellowing

Papers containing wood show signs of yellowing under the influence of light or heat [174,
Chapter 3.8], [824]. The higher the wood content, the greater the yellowing. This discol-
oration limits the usability of mechanical pulp for certain bright paper grades.
6.7 Paper � 625

The cause of yellowing is the lignins that remain in the wood portion. Functional
groups in the lignin can absorb light from 300–400 nm. In addition, some aromatic rings
can also absorb short-wave light directly. As a result, new chromophores absorb in the
UV and near-visible range and appear yellowish, e. g., aryl-α-carbonyl and quinone
structures. The reaction proceeds via excited carbonyl groups as photosensitizers and
oxygen as an oxidant. Carbonyl groups form from alcoholic hydroxyl groups of phenyl-
propane units and phenolic hydroxyl groups:

The carbonyl groups can continue to react and build complex structures. In yellowing,
we distinguish between a fast and a slow phase. In the first, fast phase, simple phenols
with catechol or hydroquinine structures are converted into quinones:

Various reactions occur in the subsequent phase, such as the nonspecific oxidation of
aromatic rings with free hydroxyl groups and of benzyl alcohol groups:
626 � 6 Structure of paint systems

The resulting radicals can react with oxygen to form oxygen functions or with active hy-
drogen atoms of other organic compounds. An example of the formation of intensively
colored quinones is the following:

During this process, carbon and ether bonds in the lignin break. Besides the yellowing,
the paper structure weakens, and the paper becomes brittle. We can see here the emi-
nent significance of the wood-free paper and the removal of the wood components for
a paper-based work of art.
7 Paint systems in art
The previous sections have presented all of the building blocks of various essential paint
systems in art. In this section, we can now take a closer look at pigments, dyes, and
specific examples of binders and auxiliary materials.

7.1 Ceramics and their painting


The artistic coloring of ceramics is one of the earliest artistic expressions after cave
painting and reached a high level already in the millennia before the birth of Christ.
Therefore, we will enter the field of paint systems and binders by considering ceramic
painting first [41, 500].
In contrast to mural or panel painting, the high firing temperatures present condi-
tions that only a few inorganic and no organic pigments can withstand when painting
ceramic objects. Nonetheless, ancient artisans applied successfully two possible solu-
tions:
– ceramic painting with highly refractory pigments before firing the object
– cold painting in the cold, ready-fired state after firing

Concerning the first solution, in the case of glazed ceramics (i. e., not classical ceramics),
we can further distinguish between
– underglaze colors applied to the body before the actual glaze; they then undergo
the high-temperature firing process
– onglaze colors applied to the finished glazed objects as a low-fusing decorative glaze
and fired at considerably lower temperatures

7.1.1 Classical ceramic painting

Ceramic painting with highly refractory colors has accompanied the development of ce-
ramics since the earliest times. The pigments are applied to the raw ware and fired with
it, which means that they must withstand the necessary high temperatures of several
hundred to about 1200 ℃ undamaged so that only a limited number of inorganic pig-
ments come into question. In antiquity, these were essentially heat-stable iron and man-
ganese oxides, calcium carbonates, and clays, providing black, brown, red, and white
colors. However, due to the complex chemistry of iron oxides, the color space could only
be opened up gradually, as the development of ceramics shows, ▶Figure 7.1.

Iron-reduction technique
The earliest technique used for painting is the iron-reduction technique, which gives the
body a black coating (black glaze):

[Link]
628 � 7 Paint systems in art

Figure 7.1: The early development phases of ceramic hot painting in the Mediterranean [41, 500].

– An oxidizing atmosphere due to full firing and sufficient oxygen supply achieves
high firing temperatures necessary for sintering ceramics. Iron compounds in the
starting materials (alumina, ocher) are oxidized to red iron(III) oxide.
– In a second step, the objects are reduction-fired by reducing the oxygen supply and
adding combustible materials, thus forming black iron(II,III) compounds. The re-
sult is a consistently black-colored ceramic since the red iron(III) oxides are also
reduced.

Iron-reoxidation technique, polychrome firing


Further development of the iron-reduction technique made it possible to use the differ-
ent colors of iron oxides in different oxidation states to produce multicolored ceramics.
Therefore, the process contains a further step for this purpose:
– Steps 1 and 2 remain the same as in the iron-reduction technique.
– Upon renewed oxidizing atmosphere in the course of cooling, certain portions of
the black surface are easily attacked by oxygen, and consequently, reoxidized to
red iron(III) oxide.

Crucial to the technique’s success is precisely determining the parts to be reoxidized.


The early reoxidation technique exploits the fact that oxygen less readily attacks thicker
applied iron-rich painting slips. If the reoxidation stops quickly enough, only thin ar-
eas are oxidized and show a red color, while thick painting layers are still black or
show different shades of brown. By varying the thickness of the layers, artists were thus
able to create subtle color transitions between red and black and achieve sculptural ef-
fects.
7.1 Ceramics and their painting � 629

The matured later reoxidation technique uses paints of different chemistry and pro-
vides safe results. Potassium-rich clays serve as black paint, which sinter more readily
than the body (or paint), which should remain red. This material creates a dense mass
that is not attacked by oxygen and is impermeable. The potassium content can be selec-
tively increased by adding potash or other potassium-rich aggregates.

Manganese-black technique
The manganese-black technique uses black iron-manganese spinels to produce the black
color. This technique is younger than the iron-reduction one and requires the pres-
ence of manganese-containing starting materials such as umber, which contain man-
ganese(IV) oxide MnO2 ⋅ nH2 O. In one step, the black color and the red iron(III) oxide are
oxidatively produced since the low-valence manganese spinels formed in the heat are
stable and retain their black color.

Pigments of the hot-painting technique


Black pigments
Three main pigment types achieve black color:
– Iron oxide black consists of iron oxides, preferably magnetite Fe3 O4 , maghemite
γ-Fe2 O3 , hercynite FeAl2 O4 , and mixed crystals of these oxides, formed during the
firing process, partly with the clay of the body. Depending on the exact composition,
the black color is altered to brown, reddish-brown, or yellowish-brown by increas-
ing proportions of hematite.
Surprisingly, the brown and colorless iron compounds maghemite and hercynite
can also contribute to the black hue if spinel formation between them and magnetite
or impurities causes oxidation states II and III to be present simultaneously. An IVCT
transition FeII → FeIII then induces the color as in magnetite.
– Manganese black comprises a less common group of black iron-manganese spinels
containing manganese in oxidation states II and III: the mixed spinel (Fe, Mn)2 O3 ,
hausmannite Mn3 O4 , jacobsite MnFe2 O4 , and bixbyite Mn2 O3 .
The manganese contents are strongly fluctuating and indicate that manganese is
partly contained only as an impurity, partly also intentionally added to increase
the manganese content, and thus the color depth. The MnII and MnIII compounds
are formed from the different manganese earths by oxygen released during fir-
ing.
– Carbon occurs in the form of graphite on ancient ceramics but cannot provide
opaque painting. Finely dispersed carbon black is better suited than graphite. How-
ever, it has the disadvantage of being very sensitive to oxidation when fired, which
is why few examples of this technique exist.
630 � 7 Paint systems in art

White pigments
Only a few compounds can be used as white pigments for hot painting:
– The Kamares pottery uses talcum Mg3 (OH)2 Si4 O10 , which is transformed in heat to
protoenstatite MgSiO3 . Since the process preserves the foliated structure of the tal-
cum, the white painting layer is not very abrasion-resistant and hardly usable for
utilitarian purposes.
– Calcite CaCO3 is best used as a fine-particle carbonate such as chalk or marl (calcite
with light, low-iron clay). The carbonate’s dissociation into CaO and CO2 during fir-
ing is compensated by subsequent carbonate reformation with carbon dioxide from
the air.
– Bright clay minerals with kaolinite Al2 (OH)4 Si2 O5 decomposing to metakaolinite
Al2 O3 ⋅ 2 SiO2 form the white ground or white color on the famous vases of the Greek
period. However, since clay components have a discoloring effect, high demands are
made on the purity of kaolin clays, so this white pigment came into use relatively
late.

Red pigments
The only heat-stable red pigment of ancient ceramics is hematite α-Fe2 O3 . Iron-rich clays
thus include, by nature, red color in the base material of ceramic objects. Ocher was
often added to increase the iron content in the base material, intensifying the color. Be-
sides red ochers, yellow ochers can also be used as a starting product since the iron oxide
hydrate is converted to hematite during the firing phase.

7.1.2 Pigments of the cold-painting technique

In cold painting, decorative colors are painted on the finished ceramic only after firing.
Since high temperatures are no longer applied, all pigments for mural or panel painting
can be used; ▶Table 7.1 lists pigments of antiquity. The cold painting extends the color
palette into the critical color space of yellow, green, and blue. White is formed by the
ground, mostly a kaolin, lime, or gypsum slurry primer. Since cold painting is not cofired
and does not form a dense coating, it lacks the high resistance against environmental
influences and abrasion of hot painting.

7.1.3 Ceramic enamel and glaze colors

The limited range of ceramic hot painting and the somewhat wider range of cold paint-
ing remained almost unchanged from the invention of ceramics to modern times. In an-
tiquity, only rarely, but later more and more frequently, glazes protected ceramic prod-
ucts, which could be self-colored or allow decorative hot painting. In close connection
7.1 Ceramics and their painting � 631

Table 7.1: Pigments more commonly found in ancient ceramic cold painting [41, 440, 495].

Pigment Composition

Yellow
Goethite* α-FeOOH
Jarosite (K, Na)Fe� (SO� )� (OH)�
Red
Hematite* α-Fe� O�
Cinnabar* HgS
Madder* Organic
Green
Malachite* Cu� (OH)� CO�
Copper hydroxychlorides (atacamite)* Cu(OH)Cl, Cu� (OH)� Cl
Egyptian green* Copper silicate glass
Blue
Cobalt blue* CoAl� O�
Egyptian blue* CaCu[Si� O�� ]
White, black, brown
Calcite* CaCO�
Dolomite CaMg(CO� )�
Huntite CaMg� (CO� )�
Gypsum, anhydrite CaSO� ⋅ � H� O, CaSO�
Carbon black, vegetable black* C
Iron and manganese oxides* (Fe, Mn)(Fe, Mn)� O�
*
The pigment was also commonly used as an artists’ pigment.

with the technique of colored glass, enamel paints emerged early on, which were also
suitable for glass painting, ▶Section 7.2.3 at p. 648. Hopper [44] depicts an entertaining
and exciting history of the development of glazes and their coloring; [42, 43, 45] supply
details of color mixing.
Modern ceramic painting applies enamel colors, which form colored glazes on
the body or an existing underglaze. They consist of three components [204, keyword
“Keramische Farben”]:
– So-called frits furnish the actual glaze. The frits absorb the color bodies during melt-
ing and coat them with a firmly adhering layer after solidification. They are thus the
binder of ceramic colors and are similar in structure to silicate glass. The earliest
frits consisted of lead silicate, obtained by applying quartz powder and lead oxide.
Today, lead, lead-boron, lead-aluminum, lead-alkali, lead-earth alkali, or soda-lime
silicates are employed.
– Colored bodies that impart color to the frits or subsequent glazes (see below).
– So-called fluxes are low-melting silicate glasses that improve the glaze’s melting be-
havior and the substrate’s wettability. Today, lead-boron, alkali-boron, and alkali-
boron-aluminum silicates are most commonly in use.
632 � 7 Paint systems in art

Based on the melting point of the glaze frits, we distinguish two series of ceramic enamel
colors:
– Underglaze colors include frits melting from 900–1100 ℃. Frits form the actual glaze
of the object and obtain their coloration from colored bodies. The earliest finds of
color-glazed objects date to around 1500 BC in Egypt.
– Onglaze colors are usually based on lead-boron silicates; their frits melt from 500 ℃.
These colors are painted over a previously applied actual glaze and are used to col-
orize, not protect, the object. Their melting temperature must be low enough not
to damage the existing glaze. Due to the necessary knowledge of glaze preparation,
onglaze colors have been available only since the eighteenth century, the age of Eu-
ropean porcelain.

[Link] Glazes
As the name suggests, glazes represent another type of glass besides the every-day
household glassware. They are vitreous coatings for ceramic materials that protect the
object and provide hardness, gloss, and color. Their composition is highly variable, and
they often have lower melting temperatures than the body itself to ensure optimal ap-
plication and uniform coating of the object. Not to be confused with glazes are glassy
coatings on metals, i. e., enamels (▶p. 307). However, enamel has a different structure
due to the required adhesion to a foreign material (metal).
As the following examples show, glazes, unlike glass, are more variable and complex
in structure and differ significantly depending on the application:

Terracotta lead glaze [43, 45] A lead glaze for terracotta ceramics contains essentially lead and
SiO2 : Red lead 80 %, kaolin 10 %, quartz 10 %. Another composition with borate content is red lead
40 %, kaolin 5 %, quartz 30 %, borax 18 %, potash feldspar 8 %, CaCO3 2 %.

Architectural ceramics glaze [43, 45] Glazes for construction ceramics can consist of sodium zir-
conium silicates: Na4 SiO4 34 %, glass 11 %, ZrO2 30 %, ZnO 15 %, SiO2 5 %, kaolin 5 %.

Majolica glaze [43, 45] Majolica wares are coated with lead feldspar glazes: Lead white 47 %, ZnO
2 %, CaCO3 3 %, feldspar 18 %, kaolin 9 %, SnO2 9 %, SiO2 13 %.

Stoneware glaze [43, 45] These may be lead glazes (PbO ⋅ 2SiO2 70 %, CaCO3 5 %, feldspar 10 %,
kaolin 15 %), or borate glazes (borax 23 %, chalk 6 %, SiO2 25 %, feldspar 33 %, red lead 14 %).

Porcelain glaze [43, 45] Glazes for fine porcelain are based on aluminosilicates: SiO2 50–80 %,
Al2 O3 0–20 %, CaO 5–25 %. Another variant is feldspar-based: SiO2 13 %, feldspar 53 %, dolomite 26 %,
kaolin 6 %.

[Link] Colorants for glazes


Glaze colorants are exposed to temperatures up to 1000 ℃ during ceramic firing. Due
to the high temperatures during sintering, only a few colorant types are suitable for
ceramic colors [16, 439]:
7.1 Ceramics and their painting � 633

– cadmium sulfoselenides stabilized as inclusion pigments


– metals colloidally dispersed in the glaze matrix
– metal oxides dissolved in the glaze
– inorganic mineral phases with the addition of coloring metal cations (mixed oxides,
complex inorganic pigments CICP), including the essential color bodies based on
zirconia (ZrO2 ) and zirconium (ZrSiO4 )

Inclusion pigments, cadmium sulfoselenides


We already know them as excellent pigments for the artist. However, they are also cru-
cial ceramic pigments, as they open up the essential color range of yellow, orange, and
red. In the form employed by the painter, cadmium pigments are not suitable for use as
ceramic paints, as they are far too thermally unstable. However, inclusion in a crystalline
matrix stabilizes the pigments against the conditions of the sintering process (inclusion
pigment). The most commonly used matrix is zircon ZrSiO4 , which forms in a solid-state
reaction from ZrO2 and SiO2 with the addition of mineralization aids. In a second step,
the zircon-containing phase grows as zircon on crystals of cadmium sulfoselenide. The
zircon phase is liquid, thanks to the mineralizers. The sulfoselenide is prepared sepa-
rately:

The complex manufacturing process makes this pigment type expensive.


Like the cadmium sulfoselenides, other colorants can be stabilized by a high-
temperature resistant layer; the best-known example is zirconium iron pink ZrSiO4 ⋅
Fe2 O3 , which we will discuss in the mixed oxides section below.

Colloidal metals dispersed in the glaze


These induce color due to free electrons that perform collective oscillations in the metal
particle. With suitable dimensions of the metal particles, the oscillation can resonate
with visible light and absorb the light, which leads to color. We have described these
so-called surface plasmons in ▶Section 1.6.4. This type of coloring is not relevant to the
painter and belongs to the domain of ceramic artists and glassblowers, ▶Section [Link].
634 � 7 Paint systems in art

The most important metal for purposes of coloring is gold. It yields a deep pink, red,
or purple color, the purple gold. Such colored glass is called gold ruby glass. Elemental
gold particles are produced via a redox reaction with SnII salts:

The reaction takes place in a slurry of kaolin or clay to immediately separate the tiny
gold particles from each other and keep them colloidal. Adding AgCl results in redder
color, and adding CoO results in purple. Silver, platinum, and copper are other metals
suitable for this type of coloration.

Transition metal oxides


These are dissolved in a glassy matrix, coloring it through an LF splitting of the metal’s
d electrons. This process is the same as the ion coloring of glass, ▶Section [Link], and
provides an effortless way of tinting glazes. However, the coloring of glass fluxes with
metal oxides has some disadvantages:
– Some metal oxides decompose to release oxygen, which can cause defects in the
glaze when it escapes.
– The conditions in the kiln influence the position of all redox equilibria, among other
things, the oxidation state in which an added metal cation is present. The oxidation
state directly determines the hue.
– The color of the metal cations depends strongly on the coordination chemistry at the
metal cation. The chemical composition of the glass largely determines the coordi-
nation chemistry. The size of the anions present decides which coordination poly-
hedra can be formed around the metal cation, and thus the coordination number
at the metal cation. As ligands, the anions directly confine the ligand field strength
that takes effect at the metal cation, and thus the color achieved.
– The presence or absence of other metals influences the compounds that form in the
glass mass, i. e., which color-active metal complexes appear.
– Before unintentional color distortions occurs, the number of impurities that can be
tolerated must be determined.

In addition to the technical problems, the prediction of glaze color is difficult, specifi-
cally when several coloring metals occur together. Detailed impressions of the complex
appearances are given in [43, 45].
7.1 Ceramics and their painting � 635

Mixed metal oxides


They are suitable color bodies, since the conditions of their formation at 800–1400 ℃ are
similar to those of the sintering process, ▶Section 3.4.3. Suitable colorless host lattices
are spinel, pyrochlore, olivine, garnet, phenazite, and periclase. ▶Table 7.2 offers ex-
amples of pigments for enamel colors and mass coloration. [204, keyword “Keramische
Farben”], [206, keyword “colorants for ceramics”], [439, 1007] provide an overview. Nu-
merous details and Recipes for representing the colored bodies can be found in [42, 43].
LF transitions lend color to this type of colorant, ▶Section 3.4.3.
Colorants such as Naples yellow or cobalt blue have long been part of the artists’
palette, and modern mixed oxides have also found their ways there, such as cobalt green
PG26, spinel brown PY119, chrome tin pink PR233, or Victoria green PG51. Many other
mixed oxide colorants in the tables, such as nickel silicate green PG56, are only used in
ceramic applications.
Green color bodies are rare as always. Today in the industry, most green ceramic col-
ors are obtained by mixing yellow and blue color bodies, e. g., zirconium vanadium blue
and zirconium praseodymium yellow, or cobalt-chromium spinel with cadmium sulfide
or Naples yellow. White color bodies are achieved by adding opacifiers, i. e., substances
that do not dissolve in the glaze but become finely distributed as particles. These scat-
ter the light and, therefore, appear white. Suitable opacifiers are the high-temperature
resistant oxides Sb2 O3 , As2 O3 , CeO2 , SnO2 , TiO2 , ZrO2 , and ZrSiO4 .
A group of mixed oxides accounting for more than 50 % of the world production
of ceramic pigments is based on the host lattices zirconia ZrO2 (naturally as mineral
baddeleyite) and zircon ZrSiO4 . Problematic are the very high melting temperatures of
zirconia (2710 ℃) and zircon (2550 ℃).

Zirconia pigments
Zirconia ZrO2 , which occurs naturally as mineral baddeleyite, serves as a host lattice.
The introduction of V2 O5 leads to yellow pigments:

Small amounts of In2 O3 yield intensive yellow; other oxides vary the hue between
greenish-yellow and orange-yellow. However, trivalent cations with a compatible ionic
radius must be available apart from VV to maintain charge neutrality. Suitable candi-
dates are GaIII , InIII , or YIII . The formation of VIII is then suppressed, and pure orange-
yellow is obtained. Vanadium zirconia colors are stable in glazes and ceramic colors up
to 1350 ℃.
636 � 7 Paint systems in art

Table 7.2: Pigments in modern ceramic glazes (hot painting) [42, p. 81], [43],
[206, keyword “colorants for ceramics”], [40, 439].

Pigment Mineral Composition

Yellow
Naples yellow PY41 Pyrochlore Pb� Sb� O�
Zircon praseodymium yellow PY159* Zircon (Zr, PrIV )SiO�
Tin vanadium yellow, vanadium yellow PY158 Cassiterite (SnIV , SnII , VV )O�
Nickel titanium yellow PY53, nickel tungsten Rutile (Ti, Ni, (Sb, W))O�
titanium yellow PY189
Chromium titanium yellow PBr24* , chromium Rutile (Ti, Cr, (Sb, W))O�
tungsten titanium brown PY163*
Zircon vanadium yellow PY160, also with Baddeleyte (Zr, (V, In, Y)III , VV )O�
indium and yttrium
Zirconcadmium yellow Zircon ZrSiO� ⋅ CdS
Zirconcadmium orange Zircon ZrSiO� ⋅ Cd(S, Se)
Nickel niob titanium yellow PY161, chromium Rutile (Ti, (Ni, Cr), Nb)O�
niob titanium brown PY162
Red
Aluminum chrome pink PR230, aluminum Corundum (Al, Cr)� O� , (Al, Mn)� O�
manganese pink PR231*
Chromium tin pink, chromium tin violet PR233 Sphene, Cr� O� in CaO ⋅ SnO� ⋅ SiO� or SnO� ,
cassiterite participation of CrIV
Zircon iron pink PR232 Zircon ZrSiO� ⋅ Fe� O�
Zirconcadmium red Zircon ZrSiO� ⋅ Cd(S, Se)
Yttrium red Perovskite Y(Al, Cr)O�
Aluminum iron red Corundum (Al, Fe)� O�
Hematite PR101 Hematite Fe� O�
Aluminum chrome red PR235 Spinel Zn(Al, Cr)� O�
Zinc iron chrome brown PBr33* Spinel (Zn, Fe)(Fe, Cr)� O�
Green and blue
Cobalt zinc aluminate blue PB72* Spinel (Co, Zn)Al� O�
Cobalt blue PB28* , cobalt turquoise PB36, Spinel CoAl� O� – Co(Al, Cr)� O� – CoCr� O�
cobalt green PG26
Nickel silicate green PG56 Olivine Ni� SiO�
Victoria green PG51 Garnet � CaO ⋅ Cr� O� ⋅ � SiO�
Zircon chrome green, zircon copper blue Zircon (Zr, Cr, Cu)SiO�
Nickel chrome green, zinc chrome green Spinel (Ni, Zn)Cr� O�
Cobalt titanium green PG50 Spinel Co� TiO�
Zircon vanadium blue PB71 Zircon ZrSi�−x O� : V�⊕
x
Cobalt silicate blue PB73 Olivine Co� SiO�
Cobalt zinc silicate blue PB74 Phenazite (Co, Zn)� SiO�
Chrome green only* Corundum (Cr, Al)� O�
Gray
Tin antimony gray PBk23 Cassiterite Sb� O� ⋅ SnO�
Iron cobalt chromium black PBk27 Spinel (Co, Fe, Ni, Mn)(Cr, Fe)� O�
Titanium vanadium antimony gray PBk24 Rutile (Ti, V, Sb)O�
Cobalt pewter gray Spinel Co� SnO�
7.1 Ceramics and their painting � 637

Table 7.2 (continued)

Pigment Mineral Composition

Brown
Manganese chrom antimon titanium brown Rutile (Ti, Mn, Cr, Sb)O�
PBr40
Manganese niob titanium brown PBr37 Rutile (Ti, Mn, Nb)O�
Iron chromite brown PBr35* Spinel Fe(Fe, Cr)� O�
Zinc iron chromite brown PBr33 Spinel (Zn, Fe)(Fe, Cr)� O�
Manganese zinc chromite brown PBr39 Spinel (Zn, Mn)Cr� O�
Nickel ferrite brown Spinel NiFe� O�
Zinc ferrite brown PY119 Spinel (Zn, Fe)Fe� O�
Iron titanium brown PBk12 Spinel Fe� Ti� O�
Black
Copper chromite black PBk22 Spinel CuCr� O�
Iron cobalt chromite black PBk27 Spinel (Co, Fe)(Fe, Cr)� O�
Manganese ferrite black PBk26 Spinel (Fe, Mn)(Fe, Mn)� O�
Cobalt ferrite black PBk29 Spinel (Fe, Co)Fe� O�
Chromium nickel ferrite black PBk30 Spinel (Ni, Fe)(Cr, Fe)� O�
Cobalt nickel zircon black only* Zircon (Zr, Co, Ni)SiO�
*
Pigments which can be used in ceramic bulk coloration (e. g., architectural ceramics).

Zircon pigments
The host lattice of the zircon ZrSiO4 has gained great importance since the middle
of the last century, as it has met the needs of the construction and sanitary indus-
tries concerning high-temperature-resistant colorants [441, 442]. Through the develop-
ment of zirconium vanadium blue ZrSiO4 : V4⊕ [443], the zircon praseodymium yellow
(Zr, PrIV )SiO4 [448], and the zirconium iron pink ZrSiO4 ⋅ Fe2 O3 [449], a primary color
triplet was provided that is not only temperature-resistant but also shows very pure
colors. However, the red pigment is not a mixed oxide pigment but an inclusion pig-
ment like the ceramic cadmium sulfoselenides. Since no commercially suitable metal
produces the red color in the zirconium lattice, it was necessary to use the colorant
hematite for the primary red color. It must be sealed in an inclusion pigment in zirco-
nium silicate [441]. The high-temperature zirconium silicate protects iron oxide from
changes during firing.
All three colorants are stable up to 1350 ℃ in all types of ceramics. However, the
color depends on parameters such as the ratio of V2 O5 to NaF, the purity of the raw
material, the particle size of the materials, mixing details, and the reaction atmosphere.
The coloring power of the blue pigment depends on the amount of vanadium that can
be introduced into the color body. Adding lead to praseodymium pigments intensifies
the hue, and adding Ce2 O4 shifts the hue to orange.
The exact position and oxidation state of vanadium in the zircon vanadium blue
were controversial for a long time. Among others, the oxidation state IV and the derived
638 � 7 Paint systems in art

possible substitutions SiIV → VIV or ZrIV → VIV were discussed, which would yield the
formulas Zr(Si, VIV )O4 or (Zr, VIV )SiO4 . Today, the prevailing view is that VIV occupies
an irregular tetrahedral site called “16g” in the zircon, and the pigment is then to be
addressed as ZrSi1−x O4 : V4⊕x [444–447]. Charge neutrality is achieved by omitting one
Si4⊕ cation per V4⊕ .
Accordingly, the color is due to LF transitions in the V4⊕ cation, specifically 2 T2 → 2 E.
Due to the distortion of the tetrahedron around the 16g site, the Td symmetry is broken.
All 3d orbitals are energetically different and form two groups, 2 E (ground state) and 2 T2
(excited state), which allow the electron transition, ▶Figure 7.2. Slight differences in the
color of the produced pigments and laboratory samples, which varies between greenish
and bluish turquoise, are explained by the lower crystal field in the industrial pigment,
in which V is highly concentrated so that V4⊕ 4⊕
x clusters instead of isolated V cations are
involved.

Figure 7.2: Splitting of the 3d orbitals of V in a tetrahedral ligand field or the further splitting in a distorted
tetrahedral field of the 16g interstitial site in zirconium vanadium blue without considering terms or LS
coupling [445]. The two orbital groups allow a transition 2 T2 → 2 E.

The zircon matrix is built up from ZrO2 and SiO2 in a solid-state reaction and calcined
with vanadium, praseodymium, or iron salt to obtain the above base colors:
7.1 Ceramics and their painting � 639

Vital is the addition of mineralization aids that allow mineral formation. Alkali or alka-
line earth salts, especially NaF, are used. The processes are complex and not yet entirely
verified, but an impression of the probable process is given in ▶Figure 7.3.

Figure 7.3: Role of mineralization aids exemplified by the formation of zircon vanadium blue [442, 443].
Above 600 ℃, the alkali halides form volatile silicon halides and liquid alkali silicates. These combine with
the above 600 ℃, likewise liquid zirconium-vanadium compounds to form V-doped zircon above 750 ℃. The
role of the mineralizing agents is thus to convert the refractory zirconium and silicon oxides into liquid or
volatile form to facilitate mass transfer and mineral formation.
640 � 7 Paint systems in art

It is necessary to use pure starting materials, making the pigments more expensive. The
price is driven up by separating the small amounts of ZrO2 from zircon sand, which is
necessary to obtain pure ceramic colors. A new process forms the colorants directly from
zircon sand without a detour via zirconia. The treatment of the starting material zircon
with alkali compounds leads to alkali zirconate silicates, which are decomposed by acid
to form a zirconia-silica mixture. If so much SiO2 is added that equimolar amounts of
ZrO2 and SiO2 are present, the zircon lattice forms in subsequent calcination [442]:

Colloid metals are not acutely or chronically toxic. However, finely divided copper,
cobalt, manganese, or nickel oxides must be handled following occupational safety and
health regulations. Zirconia and zircon are considered to be noncritical. The potentially
toxic heavy metals enclosed in the mineral matrix behave differently than the free met-
als and metal salts since their bioavailability is significantly reduced in the insoluble
matrix.

7.2 Stained glass


Artistic creations around or with glass can lure us away from our actual painting sub-
ject and lead us far into craftwork, so we will only briefly examine this area since the
chemical basics of glass (▶Section 3.9) and its coloration (▶Section 3.9.1) are already fa-
miliar to us. More detailed accounts are provided by the literature [27, 28, 35, 38, 39, 122].
Artistic expressions in this context are:
– reverse glass painting, ▶Section 7.2.1 [87, 88]
– actual stained glass, ▶Section 7.2.2
– tesserae of colored glass parts

The basis is always colorless or colored glass, which underwent a long history of devel-
opment, as essential stages display [38, 122]:
7.2 Stained glass � 641

1500 BC Egypt Cu colloid staining


1200 BC Bronze Age red glass, Cu colloid
Mesopotamia Naples yellow, yellow glazes
400 BC Celtic pottery, metal colloids (Cu and Cu� O)
0 Rome Red mosaics of Cu and Au stained glass
800 Iran, Iraq Iridescent lustre glazes on ceramics
12th/13th century Northern Europe Cu colloid staining
14th century Venice Color by metal oxides (green, red: Cu; yellow, red: Fe; purple:
Mn; blue: Co)
14th century Northern Europe Green Forest Glass (potassium silicate glass)
15th century Venice Cloudy and milk glasses, opal glasses, iridescent glasses.
Mosaic glass: ZnO in lead glass (white); enamel glass: As� O�
opalescent lead glass
16th century Venice Cristallo (soda lime glass, color compensated with Mn)
17th century Northern Europe Bohemian chalk glass (potash-lime glass color-compensated
with Mn); Au colloid stain
17th century England Lead crystal glass
Germany Gold ruby glass
18th century France Rhinestones (lead glass with: Sb, Tl, metal salts)
19th/20th century Everywhere Colorations by using a variety of metals and compounds

As early as the Bronze Age and antiquity, glass artisans could produce coloring metal
colloids by precipitating metallic copper and gold in the glass. If the glass was destined
for utilitarian or jewelry purposes, it was also possible early on to refine the surface by
applying an iridescent layer (luster). These multicolored iridescent surfaces are realized
by thin refractive layers on glass or ceramics. In the Islamic Middle Ages, alkali and
alkali-lead glazes were applied to ceramic bodies. In these glazes, metal colloids (copper
and silver) with 10–15 nm particle diameters were arranged in one or more (usually two)
ordered layers. For production, copper or silver salts (sulfates, nitrates) are applied onto
the glazed ceramic and baked in a kiln with a reducing atmosphere. During the process,
the metal cations diffuse through ion exchange into the interior of the molten glass,
where they are reduced to the metal and grow into particles of the desired size:

The metal particles’ size, shape, and organization depended on the temperature control
during the firing and were very different, resulting in various color phenomena. This
technique reached the Mediterranean area (Venice) and Spain, then Europe. Even in
modern times, luster is created by finely dispersed metals, as two recipes show:
642 � 7 Paint systems in art

Luster glaze, Tiffany 1884 2 parts SnCl2 , 1 part BaCO3 , 21 part SrCO3 , 1 part Cu(NO3 )2 evaporate.
Precipitating vapor on glassy surfaces gives an iridescent surface.

Luster glaze, Tiffany 1884 Reduce metal resinate (Bi, Pb, Ag, Cu) to metal with smoke fire and melt
into glass. Results in lustre glaze.

7.2.1 Reverse glass painting

Reverse glass painting progressed with the development of decorated glassware, stained
glass windows, and stained glass [87–89]. The following overview indicates essential
stages:

2200 BC Syria Gold etching


100 BC Syria, Rome Fondi d’oro (glass pane, gold leaf with etched drawing, ground
color)
Later Silver foil, colored etching, and painting techniques with oil
paints, casein, gouache, watercolor, tusche, inks, resin paints
13th century Italy Inlay panels in reverse glass painting or gold etching
14th century Germany Glaze reverse glass painting with gold ground
15th/16th century Germany Black/sepia drawing on the gold ground or overpainted with
glazes and painterly elements (crack in distemper, surface color
in oil)
16th century Italy Reverse glass painting after famous paintings for the wall
Since the 16th century Venice Religious motifs and high baroque art
Renaissance Graphic art recedes in favor of painterly elements; gold leaf
recedes, oil paints without contour
Baroque, Rococo Urban workshops with a commercial mode of manufacture
18th/19th century Alpine region, New flourishing by folk and religious representations
France

Unlike stained glass, in which enamel or enamel paints are applied to the glass and fused
in, reverse glass painting is a cold-color technique, i. e., no melting or baking required.
The colors adhere to the glass through suitable binders and are protected by the front
glass pane and the support. Reverse glass painting works by applying paint to a piece
of glass and then viewing the image by turning the glass over and looking through the
glass at the image. It generally has a multilayer structure:

glass – (metal foil) – color layer (opaque to glazing) – (base color) – (metal foil) –
support (board, metal ground).

Several variants of reverse glass painting have developed, all following this structure
but differing in the use of the color layers and metal foils, ▶Figure 7.4.
7.2 Stained glass � 643

Figure 7.4: Top: Layered structure of the different stained glass techniques. In the case of gold, silver, and
color etching, the drawing is etched into the upper gold or silver foil or a dark color layer; a white or light
base color or metal foil shines through the drawing. In amelioration, the drawing is also buried in the
upper gold foil; a transparent color shines through the drawing and appears shiny metallic through the
backed metal foil. In eglomisé, the drawing is buried into the upper gold foil or a black color layer; a black
base color or metal foil shines through. In stained glass, transparent or opaque layers of paint become
visible according to their layered structure. Below: Appearance of the layered structure.

Color or metal etching


In the case of etching reverse glass, the glass is painted in a single color (usually with a
dark one, then called color etching) or covered with a metal foil (usually gold or silver,
then called gold or silver etching). Next, the drawing is scratched (etched) off the color
layer or metal foil with coarse and fine tools. Subsequently, the image is painted with
a contrasting color such as white or gold or backed with metal foil (gold, silver, other
metals). When viewed from the front, the background color or metal foil then shines
out from the dark base color and achieves the image effect.
The metal foil can be glued with clarified egg white, gum Arabic, soaked quince pits,
or a mixture of linseed oil, nut oil, and possibly mastic resin. The metal foil’s formation
of luster and color was discussed in ▶Section 1.6.7.
For covering large areas in the painting, opaque paints are used, e. g., oil paints,
waterproof drying paints, or poster tempera. Water-soluble paints must be protected
with a fixative. For the background color, e. g., poster paint can be used.

Amelioration
The amelioration is a metal etching technique painted with transparent color lakes and
then backed with metal foil (silver, tin, brass). The foil gives the transparent colors a
644 � 7 Paint systems in art

shiny metallic appearance. The color lakes are bound with larch turpentine, pine resin,
or mastic resin and thinned with alcohol or essential oils. Their coloring is carried out
with transparent colorants.

Eglomisé
The glass is covered with metal foil or opaque paint, and the etched design is made
visible by dark background colors or backed with metal to lend luster to it.

Colored reverse glass painting


The reverse glass painting is an actual painterly technique. It consists of a black outline
drawing, executed as a contour with fine or broad lines. On the contour, the color layers
of the painting follow one after the other, which can be applied in glazing, semi-covering,
or covering manner. The coverage by the glass leads to a luminous color effect.
The time-saving water-oil technique was frequently used in history; for the out-
line drawing, water-soluble inks are applied to which ox gall can be added to improve
adhesion to greasy glass surfaces. Sometimes a primer with gelatin or egg white was
used. A variety of colors is applicable: acrylic paint, casein paint, opaque dispersions of
synthetic resins, gouache, tempera, ink, or tusche. In watercolors, the respective con-
temporary pigments were rubbed with gum Arabic or isinglass glue as a binder. Ex-
amples of such pigments are madder lake, gum gutter, verdigris, and Prussian blue
(Bavarian Forest, Silesia, eighteenth century) and white lead, red lead, vermilion, Prus-
sian blue, ocher, brown earth, sap green, and pine-chip soot (Austria, nineteenth cen-
tury) [87].
Oil paints were preferred for the surface color and painterly elements, which can be
diluted with linseed oil, nut oil, and turpentine oil. A fast-drying binder is boiled linseed
oil (linseed oil varnish), which simultaneously has good adhesion to glass.

7.2.2 Stained glass windows

The impressive stained-glass windows of medieval cathedrals and public buildings


prove that painting on and with glass was increasingly appreciated as a form of art in
its own right. Historical landmarks of the emergence of painterly elements on colorless
or monochromatic glass in Europe can be found in the next section, dedicated to the
actual stained glass.
The initial sacral and later public-civic stained glass windows are based on the
available solid-colored glass described in the introduction to this chapter and ▶Sec-
tion 3.9. The huge windows of gothic cathedrals consist of greenish wood ash glass (for-
est glass). However, with its high potassium content, this glass, unfortunately, is unstable
and causes significant restoration problems today. The more transparent soda glass was
7.2 Stained glass � 645

used only in cases in which the greenish tint of the glass impaired a hue. In the fifteenth
century, forest glass was further developed into wood ash-lime glass and soda-lime glass
with increased proportions of lime and soda, which was made possible by improved
furnace technology and change of composition. The melting temperature could be in-
creased from about 1100 ℃ to 1300 ℃, resulting in a more homogeneous, purer, thinner,
and more transparent glass. With the help of refining agents such as arsenic and anti-
mony compounds apart from the already known manganese ones, glassworks were also
able to increase the clarity of the glass. These factors combined produce a material with
a previously unknown clarity and light transmission [27, 28, 86].
For the production of medieval glass windows, the finished panel glass was stained
according to state of the art by metal cations and metal colloids as described in ▶Sec-
tion 3.9.3. We often find the ingredients mentioned in ▶Table 7.3, but all agents men-
tioned in ▶Table 3.11 except calcium antimonate were employed for glass staining.

Table 7.3: The most important coloring components of medieval church glass (stained glass win-
dows) [28, Chapter 10], [621]. In addition the coloring agents described in ▶Section 3.9.3 were used.

Color Components

Emerald green Wood ash lead glass with copper oxide (CuO) in oxidizing atmosphere. Later
soda lime glass with CuO, FeO, and Fe� O�
Red Wood ash glass (later soda lime glass) with copper in reducing atmosphere,
gives Cu� O (pigment) or Cu� (colloid)
Blue CoO in soda ash glass or soda lime glass with lead (as wood ash glass is too
greenish). Later soda lime glass with FeO and MnO, turquoise with CuO and iron
Yellow Iron polysulfides (“carbon amber”) in wood ash glass, produced from
sulfur/carbon. Later soda lime glass with colloidal silver and (for brown yellow)
MnII /FeIII
Purple, incarnate MnIII in wood ash glass

The following steps follow the preparation of the colored panel glass:
– Cut (solid-colored or colorless) glass panels into pieces according to the drawing.
– Break glass pieces cleanly into shape.
– If necessary, paint with black and silver solder (see next section), and bake the color
onto the painted glass in the furnace.
– Apply lead, i. e., set the finished glass pieces in lead tape.
– Solder the lead bands.
– Cement the edges of the lead bands with a mastic made of chalk and linseed oil
[37].
646 � 7 Paint systems in art

7.2.3 Stained glass

The actual, painterly-decorative glass work is as old as the material itself, as we see in
the following overview [122], [27, Chapter 7], [28, 86, 620].
After decorating glass and ceramic objects with enamels in antiquity, a phase of
decorating and later painting on glass vessels (hollow glass painting) followed in the
Roman period. The colorations obtained in this process are based on the ion and an-
nealing coloration mechanisms mentioned in ▶Section 3.9.1. In the early Middle Ages,
a painterly technique became established, associated with the mosaic-like picture design
by monochromatic glass pieces. The sober austerity is accompanied by graphic elements
such as dark opaque contours and hatchings, leading through a more unrestrained ap-
plication to semiopaque or glazed layers and tonal modeling with high plasticity. Finally,
the strict graphics recedes in favor of free painting. Stepping out of the sacred frame-
work, panel painting on glass develops.

Black solder drawings, silver solder, sanguine, or iron red solder


Painting in the strict sense of the word on flat surfaces was initially done in the early
Middle Ages on colorless or polychrome glass with black solder paint. It was prepared
from a copper and iron oxide mixture, a low-melting colored lead glass, water, vinegar
or turpentine oil, and a binder such as gum Arabic. The black solder, thickly rubbed
with turpentine oil, serves to draw the outline, the contours of the picture, and graphic
elements such as hatching. Aqueous black solder allows the colored glass to be painted
thinly over the entire surface and softens the luminosity of the glass in this area. Bristly
or fine brushes can add structures to the brownish-greyish paint, and highlights can be
brought out again. The artist can draw delicate structures, facial features, and folds in
robes and objects with thicker black solder.
After the painting is finished, the glass panel and its black solder drawing are
fired at about 600 ℃. During this process, the glass components of the solder melt and
firmly bond the metal oxides to the glass panel. However, the temperature is not so high
that the metal oxides dissolve in the glass ionically but instead form a transparent to
opaque brown-gray mass that helps attenuate transmitted light according to the painted
structure.

Black solder, after Theophilus [27, Chapter 7.6] Copper oxide, green glass, “sapphire-colored
glass” (probably lead glass), wine, or urine

Black solder [28, Chapter 10.3] Iron oxide, copper oxide, lead glass

Black solder [36] Iron or copper oxide, lead glass, vinegar or turpentine oil, gum Arabic

The role of vinegar is explained by the fact that lead(II) oxide in lead glass is readily
attacked by acetic acid at higher temperatures. Consequently, the lead glass becomes
etched and porous; the metal oxides can penetrate the glass more easily [620].
7.2 Stained glass � 647

Predynastic Ancient Green and blue glazes on quartz pottery with CuII . Yellow opacities with
Egypt, Pb/Sb and Pb/Sn, white ones with Ca� Sb� O� , black ones with FeIII , purple
Mesopotamia ones with MnIII , red ones with Fe� O�
1470 BC Egypt Yellow enamel, yellow glass (Naples yellow)
500 BC Babylon Yellow enamel tiles
6–8 century Egypt Silver yellow
841 Painting on glass
8th/9th Europe Colored windows in Carolingian and Longobard churches and palaces; mainly
century green, besides yellow, blue, and brown. Black solder for drawing and painting
9th century Iraq, Middle Colorfully decorated glass with enamel paints, luster enamel; silver yellow
East
12th century Europe Beginning of forest glass production (wood ash glass)
Around 1250 Europe Need for large-scale stained glass due to the new construction of Gothic
cathedrals, e. g., at Cologne, Regensburg, Strasbourg, Freiburg
13th century Europe Early Gothic black solder painting, 1134 grisaille painting in Cistercian
monasteries
12th to 14th Syria Colorfully decorated jars with enamel paints, chandelier enamel
century
13th/14th Europe High phase of glazing of Gothic cathedrals
century
1300 Europe Reddish-brown paint for contours and incarnate
1300 Europe Silver solder for yellow glass
14th century Europa Etching in black solder, modeling with brushes (stippling). Black solder
residue clouds the colored glass
15th century Europe Panel painters as stained glass artists; the emergence of landscape depictions
15th century Venice Enamel with opaque colors on low-iron, colorless glass
15th century Venice Faience, opaque paint on pewter white glaze, enamel painting on glass,
opaque paint on colored glass
15th/16th Europe Harder and purer wood ash lime glass
century
15th century Europe White and yellow enamel paint for tinting and shading, red glazes (sanguine
or iron red) from Fe� O� , glazing to opaque, brown to yellow, reddish-brown
to flesh-colored
15th century Europe Fine drawing, hatching, and similar graphic techniques provide chiaroscuro,
tonal values, plasticity. In the late Gothic more painterly treatment, lead rods
recede, black and silver solder with etching and etching techniques. Enamel
or enamel colors on colorless or monochrome glass
16th century Venice Enamel painting on clear glass, opaque color
Germany, Cold painting on glass, fine painting on glass with transparent colors (black
Austria, solder, silver solder, iron red, manganese violet, first onglaze colors for
Venice porcelain)
16th century Europe Residual scratching out the paint with metal brushes (scratch dabs) results in
brilliant highlights
17th century Venice Milk glass and iron red (sanguine)
648 � 7 Paint systems in art

Silver solder, which appeared later in Europe, has a transparent character. It consists
of silver chloride and sulfide, possibly Sb2 S3 , ocher and clay, water, and a binder such
as gum Arabic or oil. As soon as it is applied (usually covering a large area) to the glass
panel, it is fired at about 600 ℃. During this process, colloidal silver forms by reduction
with FeII ions. As a result, the glass becomes transparent and luminous and bright lemon
yellow to yellow-orange, depending on temperature control and firing time.

Silver solder [27, Chapter 7.7] Silver powder, antimony(III) sulfide, clay, or ocher or loam, water,
sticky substances

For painting incarnate served the iron red or sanguine of Fe2 O3 , with which red glazes
could be achieved. Depending on the composition and the conditions of application and
firing, the glass painters could paint glazing to opaque layers and colors from brown to
yellow-brown and red-brown to a nude color.

Enamel paints
The developing enamel painting resides on enamel paints, which are very similar
to the enamels and ceramic paints mentioned in ▶Section 3.10 and ▶Section 7.1.3
[27, Chapter 7], [28, 29]. Some ancient Egyptian-Mesopotamian works from predynastic
times show green and blue glazes on quartz ceramics (faiences) colored with copper.
Yellow opacities with Pb−Sb and Pb−Sn, white ones with Ca2 Sb2 O7 , black ones with
FeIII , purple ones with MnIII , red ones with Fe2 O3 , and the colorants mentioned in ▶Sec-
tion 3.9.3 complete the palette [28, Chapter 2–4]. Actual enamel paints appeared in the
High Middle Ages in the Islamic sphere of influence and are based on halophyte ash
lead glass for white, yellow, green, and black, and halophyte ash glass for blue and
red respectively. Twelfth-century enamel paints from the Middle East, e. g., show the
following compositions: Blue CoII , black FeII,III , white SnO2 , yellow PbSnO3 , and green
CuII [28, Chapter 9.7]. Medieval glazes on glass from the Islamic region around 1300
show the composition mentioned in ▶Table 7.4.

Table 7.4: Composition of medieval glazes from the Islamic region, around 1300 [29].

Colorless base glass Halophyte ash soda lime glass, MnO� for color compensation
White Soda glass, opacified with SnO� and Ca� (PO� )�
Blue White enamel, stained with lapis lazuli or CoII
Red Soda lime glass, Fe� O� as hematite, or lead glass with Fe� O�
Pink White enamel and hematite
Purple Blue and red colors with lapis lazuli and hematite
Yellow Lead glass, lead-tin yellow
Green Lead glass, lead-tin yellow or lead antimonate, copper oxide
Brown Lead glass, chromite (Mg−Fe−Al chromium oxide)
Black Lead glass, chromite (Mg−Fe−Al chromium oxide), copper
7.3 Fresco (mural painting) � 649

Enamel paints are mixtures of easily melting glass (flux), usually lead glass and coloring
metal oxides or metals mixed in water and gum Arabic or oil and turpentine to form the
paint. The paintability improves by adding egg, honey, or sugar syrup. Binders and ad-
ditives are not critical since they burn when heated. The additives listed in ▶Table 3.11
were already used for glass coloring in historical enamel paints. As more and more lay-
ers of paint were superimposed with the development of painterly technique, aqueous
and oily painting media were often used alternately so that the following layer of paint
could not wholly or partly dissolve existing ones. The final fixation of the color took
place only at the end of the firing.
At about 600–700 ℃, the enamel color fuses by melting the glass with the base glass.
White mosaic glass (▶Section 3.9) was used as the basis for the initially opaque colors,
and transparent colors were obtained by a lead glass containing boric acid. In contrast
to enamel colors for metal coating, glass enamel colors contain less alkali to adapt the
thermal expansion to the base glass.

7.3 Fresco (mural painting)

In addition to ceramic painting, the fresco is one of the oldest techniques already exe-
cuted in antiquity, ▶Figure 7.5 [121, 128, 129, 131, 132].

Figure 7.5: Schematic temporal use of the different techniques of mural painting [120, p. 195].
650 � 7 Paint systems in art

From ancient literature to modern linguistic usage, all kinds of mural paintings are of-
ten called frescos. In addition, there is the so-called lime painting. Therefore, we must
conceptually distinguish the genuine fresco technique (fresco buono) from fresco secco.
The fresco technique requires that the layers of paint are applied to a wet (freshly laid)
lime plaster, bond with it, and dry together to form a homogeneous layer. The secco tech-
niques are distinguished from this, in which the paint is applied to dried plaster. From
ancient times until today, paints bonded with gum, glue, or tempera paints were used
for the secco technique. Many mural paintings from the eleventh century are bound in
the milk of lime, hide glue, casein, or linseed oil [54, 121, 509–513].

7.3.1 Fresco-buono technique

The genuine fresco technique is ancient [120, p. 195]. Already 1500 BC excellent mural
paintings were executed as fresco buono by the Minoan culture on Crete, some of which
show surrealistic-looking undersea landscapes. The technique was widespread in the
Aegean and was widely used by Etruscans and Romans until the fourth century. It contin-
ued in Byzantium in late antiquity and later reexported to Italy. The Trecento extended it
with secco techniques, making possible a sophisticated painting technique with a higher
degree of realism. In the Southern Alpine region, this tradition continued until modern
times. In the Northern Alps, the buono technique does not have a tradition; it was first
introduced in the Romanesque period, culminating in the illusionistic ceiling paintings
of the Baroque period. The rediscovery of the true fresco occurred in the nineteenth
century by the Nazarenes.
In contrast to the practical execution, which requires high craftsmanship from the
painter, the chemistry of fresco buono is comparatively simple [121, 944]. It comprises
two steps; the first is carried out industrially in producing lime mortar. It produces cal-
cium oxide (burnt lime) through burning limestone (calcium carbonate) at high temper-
atures. Subsequently, the calcium oxide is turned into calcium hydroxide (slaked lime)
by immersion in water:

The actual painting process begins with the painter or plasterer who uses sand and
slaked lime to create a layer of plaster that serves as a base for painting, and at the
same time, as a brilliant white color. The painter mixes the pigments with water or lime
milk and paints wet on the wet plaster. Since the plaster is wet during painting, water or
7.3 Fresco (mural painting) � 651

lime milk combines effortlessly with its moisture, and the applied pigments sink into it.
There they are enclosed by lime solution. During drying, the slaked lime sets by absorb-
ing carbon dioxide from the air under recarbonatation, in which silicate components of
the sand also participate:

The pigments are enclosed in a solid but thin, translucent lime layer. This protection
against the outside world explains the stability of this type of painting. The high speed
at which the lime sets is somewhat tricky for practical work. Fresco painters, therefore,
divide the mural surface into “giornata” (Italian for “days”), i. e., smaller sections that
can be plastered and painted within a day.
If the setting process is already advanced or even completed, the painting can no
longer be executed al fresco. The only ways out are to paint in secco technique on the dry
plaster, continue in lime paint, or knock off the hardened layer of plaster and replaster
the section.

7.3.2 Lime-painting technique

Besides the fresco buono technique, since 1500 BC, the so-called lime painting spread in
the Aegean, Rome, Byzantium, and the Southern Alps area [120, p. 195]. In Asia Minor
and north of the Alps, it was a mural painting technique used in all epochs. Until the
Romanesque epoch, it was even the only one known north of the Alps.
In contrast to buono painting, the plaster substrate for lime painting is moist or even
dry, i. e., already set. A whitewash of slaked lime is then applied, in which the colors can
then be painted wet-on-wet. Like the al fresco technique, the colors are bound by the
setting lime, enclosing the pigments with a layer of lime. Thus, the bonding is much
better than the secco techniques, where the paint adheres superficially to the plaster
but less than the fresco buono technique.
The appearance of a lime painting is duller, more graphic, and two-dimensional than
a fresco, which acquires a subtle shimmer and depth through the alignment of pigments
and lime crystals.

7.3.3 Fresco-secco technique

Painting al secco is any painting executed on dry plaster. Depending on the binder, there
are different variants:
652 � 7 Paint systems in art

– Binding agents of glue or casein painting is animal or vegetable glue from, e. g., pro-
tein, animal hide, fish bones, casein, starch, or gum Arabic. This technique is attested
already around 3300 BC in Egypt.
– In tempera painting, an emulsion is used as a binder. Water-soluble components are
casein, animal glue, or plant gums. Water-insoluble components are oils, resins, or
waxes. The origin of tempera is assumed in the northern Europe of the thirteenth
century, in England or northern France. From there, it spread to the south but could
not displace the fresco buono technique in Italy.

7.3.4 Mixed techniques

In mixed media paintings, the artist paints more or less extensive sections of a fresco al
secco. This approach may be necessary to make corrections, add finishing touches and
details, or save time. However, a painting is only expressly denoted as executed in mixed
technique if the painter has performed a considerable part of the fresco al secco or has
added the secco technique for artistic intent.

7.3.5 Pigment degradations

Pigments can change in undesirable ways due to their chemistry or environmental in-
teractions. Especially mural paintings, which often date back to the Middle Ages, late
antiquity, or even the early period, i. e., 400–4000 years of age, show an abundance of
damage patterns that can be traced back to such unwanted pigment degradations. Some
of these are:
– total loss of coloration, leaving gray or black colors
– browning of red lead or lead-tin yellow
– blackening of lead white or vermilion
– greening of blue copper pigments
– loss of yellow, purple, and blue colors in mandorlas or rainbow representations

Degradations also affect drawings, watercolors, and panel paintings, although usually
not to the same degree as mural paintings, ▶Section 7.4.11. Unfortunately, the level of
knowledge is still low; only a few systematic studies of a chemical nature are available.
According to previous research, pigments containing copper, lead, and mercury, are par-
ticularly affected [481].
Since antiquity, there has been and still is a discussion about whether certain
pigments are generally unstable in lime binders or intrinsically unstable or whether
weather conditions, moisture, and atmospheric gases cause degradations. Pliny the El-
der declares, e. g., azurite, orpiment, realgar, cinnabar, red lead, lead white, Stil de grain,
indigo, carmine, and madder lake to be unsuitable pigments for fresco. Knoepfli [121]
explains that it was not the lime sensitivity of azurite, which posed the problem, but the
7.3 Fresco (mural painting) � 653

introduction of chloride ions. Also, the transformation of azurite into malachite had not
been proven. Lead white would become brown by oxidation or by sulfide formation,
but not due to lime intolerance. Furthermore, cinnabar was sensitive to light and heat,
not to lime. On the contrary, lake pigments and specific copper, lead, mercury, or arsenic
pigments were especially sensitive to lime.
The situation is further complicated because, over the millennia, a large number of
different names circulated for the same pigments, or the same names denoted different
pigments, names had fallen into disuse, or the chemistry of the designated pigments was
or is wholly or partly unclear. Earlier restorers often only described the changes visually
without chemical analyses of the original pigments, the degradation products, and the
environment.
[481] systematizes the degradations according to cause, among other things:
– The production phase lays the foundations for the durability of a pigment in the
paint system. Future events depend on the purity of the raw pigments and the ad-
ditives used in their production.
– The painting phase influences the pigment through the presence of other pigments
and the type of inclusion or separation in the painting layer. This structure of the
painting layer enables or prevents reactions of the pigments with each other and
with environmental conditions such as atmospheric gases.
– During the weathering phase, the pigments are exposed to various salts (chloride,
sulfate) from the soil and masonry, moisture, and atmospheric gases such as CO2 ,
SO2 , or O3 . Lime can thus transform into gypsum and coat the pigments with gypsum
crusts.

These few examples should illustrate that the chemistry of pigment degradations on
walls or panel paintings is complex, so we can only depict exemplary processes here.

Reactions of copper pigments


Blue copper pigments can turn green and degrade to complete blackening:

Blue copper carbonates convert to green copper chlorides by introducing chloride ions
in this process. The necessary chloride ions may come from a red lead, in which chloride
was always present due to the production process. In industrial time, factories releasing
Cl2 - and HCl-containing gases are a regional chloride source. Under the influence of an
oxidizing agent, copper pigments can turn into black oxide in the final stage of decay.
Conversely, green copper pigments can turn blue:
654 � 7 Paint systems in art

The sulfate anion necessary to form the blue, basic copper sulfates can be derived from
gypsum moieties or SO2 from the air. Green copper pigments may also blacken as a result
of oxide formation:

Reactions of lead pigments


In the case of the lead-containing pigments, mainly changes in the orange-red red lead
used until the late Middle Ages were investigated. Phenomena observed are the brown-
ing of red lead, which is typical of Gothic paintings, blackening, or fading:

The black or, when mixed with red pigment, brown color impression is caused by black
lead dioxide, while a decolorization is caused by transformation into white or colorless
lead salts. Lead dioxide also causes the blackening of lead-tin yellow, and the formation
of colorless sulfate is responsible for the decolorization of massicot:
7.4 Oil paint � 655

Reactions of vermilion
Vermilion can undergo blackening by converting the trigonal red to the cubic black mod-
ification. The phase transformation proceeds under the action of light and heat:

Subsequent reactions can lead to colorless lead salts, e. g., mercury(I) chloride.

7.4 Oil paint


Oil painting as we know it dates back to the late Middle Ages and has become the stan-
dard technique for European panel painting since about 1500. This success is undoubt-
edly due to several characteristics that other paint systems did not possess or came to
possess only in modern times:
– The slow drying of the oil implies a long open time of the paint, i. e., the image re-
mains moist and changeable for a long time.
– Fine color gradations are possible by smudging and blending. The earlier tempera
technique could only achieve this effect by stroking.
– Oil paint can be applied both thinly glazed and plastically thick.
– The colors do not change during drying.
– The physical film shows the clear surface structure (brushwork) and promotes deep
light (pure, intensive colors) instead of scattered light (matte chalky colors).

7.4.1 Basic composition of oil paints

Oil paints consist of pigments embedded in a drying oil, usually from linseed. Therefore,
we can rub simple oil paints according to this basic recipe:

Oil paint [74, Chapter 9], [75, p. 178] 1–3 parts of pigment, 1 part linseed oil or another drying oil
(for water-dilutable paints, modified oil together with emulsifier), possibly fillers, and siccatives;
for resin-oil paints, a resin (dammar, mastic, acrylic resin). Keep the amount of oil at a minimum; to
be more precise, tabulated values of the oil absorption indicate how much oil is required for each
pigment; see [75, p. 65ff] for examples.
656 � 7 Paint systems in art

Mix pigment and binder with a flexible palette knife or spatula to a stiff paste. Ground the paste
with a muller. Work in more pigment if suitable.
Be careful not to inhale hazardous or toxic pigments and avoid direct body contact!

The literature cited, which deals explicitly with self-producing oil and other paints, give
detailed practical tips and procedures.
Oils suitable for oil painting are drying fatty oils (see the next section), which dry due
to their unsaturated character if oxygen is added with cross-linking of the oil molecules
and form a transparent, viscous mass, ▶Section 7.4.3. Fillers are included in expensive
or strongly coloring pigments. White pigments or cheap calcium carbonates can lighten
the shade or stretch the pigment. Siccatives support and accelerate the drying process
by their catalytic action, ▶Section 7.4.5; they are compounds with metal as an active
component, e. g., cobalt or zinc. Some pigments, such as cobalt blue, already contain
these elements. In addition to their color effect, they also have an accelerating effect on
drying.
Grinding oil paint is not a simple process. At the first attempt, novices in painting
produce a paste that resembles oil paint and can be picked up with brushes but crumbles
and becomes dusty when applied to the canvas. The reason lies in the energetic barriers
that occur during the interaction of pigment particles and the binder, which we have to
overcome mechanically or with auxiliaries. ▶Section 6.3 depicts the process of grinding,
[74, Chapter 9], [75, p. 187] explain the procedure for homemade paints.

7.4.2 Types of oils

Drying fatty oils are responsible for the properties that characterize oil painting: The
formation of a durable transparent film with embedded pigments allows us to work
with surface light and deep light. Reason enough to take a closer look at these crucial
substances [280, p. 581ff].
We can divide fatty oils into nondrying and drying oils. Nondrying oils do not poly-
merize but cannot dry by evaporation due to their high molecular weight. Therefore,
they are of no importance as painting material and would even destroy the paintings
due to their persistent softness. The so-called semidrying oils such as soybean or sun-
flower oil, highly esteemed in the kitchen, also dry too slowly or insufficiently to be
considered painting material. However, the drying oils are distinguished because they
remain liquid after application, allow to work a la prima, then dry chemically, i. e., poly-
merize, and form a film. The time required for drying depends on the type of drying oil
used. In sum, these oils are the fundamental substances for the oil painting technique.

[Link] Drying fatty oils


Drying oils are obtained by pressing plant seeds or nuts. The most significant ones are
linseed oil (from seeds of the linseed plant, whose fibers provide us with linen), poppy
7.4 Oil paint � 657

Table 7.5: Composition of essential vegetable oils in percent [280, p. 589]. The oils commonly used in the
art environment are highlighted in bold type.

Fatty acid Canola Wheat Maize Linseed Poppy Walnut Safflower Soy Sunflower
germ germ seed

16:0 4 17 10.5 6.5 9.5 8 6 10 6.5


(palmitic acid)
18:0 1.5 1 2.5 3.5 2 2 2.6 5 5
(stearic acid)
20:0 0.5 0 0.5 0 0 1 0.5 0.5 0.5
(arachidic acid)
18:1 (Δ� ) 63 20 32.5 18 10.5 16 12 21 23
(oleic acid)
18:2 (�,�� ) 20 52 52 14 76 59 78 53 63
(linoleic acid)
18:3 (�,��,�� ) 9 10 1 58 1 12 0.5 8 0.5
(linolenic acid)

seed oil, safflower oil, and walnut oil. As a glance at ▶Table 7.5 confirms all of these oils
contain high amounts of polyunsaturated compounds responsible for polymerization.
They are bound in the oil as esters with glycerol:

The table shows that the classic linseed oil has the highest content of triple unsaturated
linolenic acid. As we shall see in a moment, all reactions leading to oil drying begin ear-
lier and proceed faster with an increasing number of allyl groups. Linseed oil, therefore,
has particularly pronounced drying properties.
In addition to the ability to polymerize, the double bonds add a further property to
the molecule, namely the lowering of the melting point. Since the double bonds have a
(Z)-configuration, the fatty acid molecules become more curved as the number of double
658 � 7 Paint systems in art

bonds increases. This shape prevents the formation of crystalline regions in the molec-
ular structure. Oils are therefore liquid, while fats, which consist mainly of saturated or
less unsaturated fatty acids, are solid or semisolid (→ fat hardening by hydrogenation
for margarine production).

[Link] Linseed oil variants


Besides raw linseed oil, several variants are manufactured and offered for commercial
and artistic use [74, p. 75], [75, p. 169]. Each exhibits particular properties suitable for
different fields of application in oil painting, ▶Table 7.6.

Table 7.6: Variants of raw and treated linseed oil, in order of decreasing superiority and suitability for oil
painting [75, p. 169].

Oil Treated oil Refined oil

Linseed oil, cold-pressed Stand oil Alkali-refined oil, varnish oil


Binder - Binder, varnish
Grinding and paint vehicle - Grinding and paint vehicle
Paint medium Glazing and leveling paint medium Paint medium
Yellowing, slow-drying nonyellowing, fast-drying Color-stable
Sun-refined or bleached oil
Varnish ingredient
Leveling paint medium
Bleached, fast-drying
Linseed oil, hot-pressed Blown oil, boiled oil
not recommended not recommended

Cold and hot-pressed linseed oil


Linseed oil can be obtained in several ways. Cold pressing provides the traditional, very
pure linseed oil [74, p. 75], [75, p. 169]. After some aging time, mucilaginous sediment set-
tles out, and the oil is ready for use after filtering. It is valued as the highest-quality oil
for painting.
However, the oil yield increases through extraction with hot water vapor, which de-
livers an oil contaminated with steam and water-soluble substances. Therefore, it must
first be purified or “refined” before being suitable for artistic use and is regarded infe-
rior to cold-pressed oil before purification.

Refined linseed oil, varnish linseed oil


Hot-pressed linseed oil must be purified (“refined”) before it is applicable for oil paint-
ing [74, p. 75], [75, p. 171]. Commercially, the oil is mixed with sulfuric acid, water, and,
7.4 Oil paint � 659

eventually, bleaching agents. The treatment removes mucilaginous and yellowing com-
ponents. Subsequently, all traces of acid and water are removed. More modern treat-
ment employs superheated steam instead of chemicals; alternatively, alkali instead of
acid can be used to purify the oil, yielding alkali-refined oils with a low acid number,
suitable for varnishes and painting.
Varnish linseed oil denotes an alkali-refined linseed oil that “do not break”, i. e., oil
that can be heated rapidly to 260 °C without flocculating. This type of oil is suitable for
the production of clear varnishes due to its color stability, widely available and can be
employed instead of cold-pressed linseed oil for general application. Varnish linseed oil
is not to be disturbed with linseed oil varnish, ▶Section 7.4.6.

Stand oil
Stand oil is linseed oil, heated to 275–300 °C for several hours under strict exclusion of
air or oxygen [74, p. 75], [75, p. 172], [185, Volume 1]. The result is a heavy, viscous oil with
a viscosity of about that of honey. For use, it is thinned with several parts of turpentine.
Stand oil is less suitable as an actual binder (although it has not lost its binding
power) or as a vehicle for grinding oil paint. However, artists highly value it as a sig-
nificant additive to glazing and painting mediums, varnishes, oil paints, and tempera
emulsions, increasing viscosity and gloss and introducing pleasant flowing properties.
Its principal characteristic is its capability of leveling, i. e., to dry to a uniform, smooth,
enamel-like film free of brush marks. It imparts this property also to mediums and paints
when added. Simultaneously, its wetting power and dispersive properties decrease, and
drying speed, leveling, and protective properties increase.
While the heating process was carried out in open kettles in former times, the re-
sulting product was inferior due to oxygen access. Nowadays, high-quality stand oil is
manufactured industrially, using specialized equipment and high vacuum or a CO2 at-
mosphere. Lighter and heavier grades of viscosity exist, but there is no particular target
viscosity.
Stand oil is partially polymerized but not oxidized. Diels–Alder reactions be-
tween different triglycerides form the polymers, cross-linked via carbon bridges, ▶Sec-
tion 7.4.4. Especially drying oils with conjugated double bonds polymerize rapidly at
high temperatures since these already have a diene structure.

Blown oil
Blown linseed oil is a result of blowing air at moderate temperature through linseed oil or
other drying oils [74, p. 76], [75, p. 173]. The high oxygen supply accelerates the processes
of oxidative drying, and the oil polymerizes as it does when exposed to air after the
painting is finished. The thickened, viscous oil is prepolymerized and cross-linked by
ether and peroxo bridges. Since cross-linking has already taken place, the blown oil no
longer has any binding power and is inferior to raw linseed oil or stand oil. It is not
660 � 7 Paint systems in art

recommended for permanent painting, and painters hardly ever use it. Instead, it is
intended for coatings and commercial paints.

Sun-refined or bleached oil


In the Renaissance, a rapidly drying, bleached oil was made by frequently and vigor-
ously shaking linseed oil with water and exposing this mixture to intense sunlight for
several weeks while allowing the access of air [74, p. 75], [75, p. 173]. The length and in-
tensity of this treatment were manifold and not concise. Before use, the resulting oil
was filtered to remove gelatinous impurities and separated from the water. By this treat-
ment, the linseed oil becomes thick and bleached, partially oxidizes, and polymerizes by
cross-linking with oxygen bridges. It is an intermediate state between stand and blown
oil, in which the processes of regular drying are partially anticipated and, therefore,
sometimes regarded as similar in quality to blown oil. Nevertheless, it was used with
good results in the past. It is suitable for clear varnishes, glazes, and painting medium
purposes.

Boiled oil
Boiled linseed oil is linseed oil heated with siccatives until it achieves a slightly thickened
consistency [74, p. 76], [75, p. 173]. It is not recommended for permanent painting and
intended for coatings and commercial paints.

7.4.3 Drying of oils, film formation

One of the most striking phenomena of drying oils is their ability to form a transparent,
solid film, which embeds and protects the pigments. For this film formation to take place,
we must expose the fresh oil painting to both oxygen and light.
To understand what happens during this process and why usually harmful envi-
ronmental parameters of light and air are necessary, we need to look at several inter-
connected processes [189, 293, 294], [280, p. 175ff], [892–904]. Many decades of chemical
research have shown that hydroperoxides formed from fatty acids by oxygen addition
play a prominent role in these processes. Several phases of drying group the processes:
– Induction phase. Since all drying reactions are by nature radical, the inherently
present antioxidants first appear. Only when their capacity is exhausted, the actual
drying begins.
– Formation of the hydroperoxides. This phase provides the reactive species from the
all-cis acids under cis-trans isomerization and thus determines the rate of oil drying.
– Decomposition of hydroperoxides. The formed hydroperoxides decompose rela-
tively readily during the drying phase, which takes weeks to months. The decom-
position leads to two types of reactions:
7.4 Oil paint � 661

– Cross-linking. The radicals formed during hydroperoxide decomposition cross-


link molecules and lead to the desired film formation by building a carbon-
oxygen network.
– Cleavage. The hydroperoxides decompose while the oil components undergo
cleaving. As the cross-linking and molecular size is reduced in the process, the
quality of the film decreases.
– Hydrolysis of the glycerides. In the months and years after drying, the glycerol esters
react to moisture and form glycerol and free acids by hydrolysis. Since hydroper-
oxides are always formed by atmospheric oxygen, slow cleavage and degradation
reactions of the oil components occur.

During the oil drying process, the proportions of certain compound classes show a typ-
ical time course: The number of double bonds decreases with time, the amount of cis
bonds steadily diminishes, and that of the trans bonds increases at first and then also
decreases. The amount of hydroperoxides increases to a maximum and then falls, while
the amounts of aldehydes, ketones, ethers, and peroxo-ethers increase. Polymeric ethers
and peroxo-ethers are the desired results of film formation, while aldehydes, ketones,
and the resulting carboxylic acids and alcohols are signs of film degradation.
Furthermore, in more detail, we will consider the basic processes shown summarily
in ▶Figure 7.6. In doing so, we will encounter the question of which of several possible
hydrogen atoms is radically abstracted. Looking at the binding energies of various types
of hydrogen [280, p. 177], we recognize that the bis-allyl hydrogen is the weakest bound
since the remaining bis-allyl radical is the most stabilized with respect to mesomerism:

Formation of radicals and hydroperoxides


The radicals at the center of all reactions are formed either directly or through hydroper-
oxides, which decompose into radicals:
– Autoxidation. The homolytic cleavage of a hydrogen atom from a fatty acid forms
an alkyl or allyl radical, which is then oxidized to the peroxo-radical and maintains
the chain reaction. This process is very slow at room temperature and occurs only
with unsaturated fatty acids.
– Lipoxygenase reaction. In living cells, the enzyme lipoxygenase generates hydroper-
oxides from fatty acids. Since all paint oils are produced from plants, even fresh oil
always contains a small number of hydroperoxides.
662 � 7 Paint systems in art

Figure 7.6: Basic reactions in the drying of oils. Initiation phase: formation of the starting radicals by au-
toxidation, lipoxygenase reaction, or photo-oxygenation. Chain reaction phase: chain reaction between
C-radical and peroxo-radical (also part of autoxidation). Cross-linking phase: formation of peroxo and ether
bridges. LOX: lipoxygenase; dashed arrows: photo-oxygenation; bold blue arrows: chain reaction; red ar-
rows: cross-linking reactions.

– Photooxygenation. Irradiation with light accelerates the oxidation of the fatty acids
and the formation of hydroperoxides by a factor of one hundred. This speed advan-
tage is the reason why we expose our paintings to light and air during drying to
promote the drying of the oil layer.

Autoxidation
The autoxidation or direct oxidation of fatty acids by the ordinary triplet oxygen
7.4 Oil paint � 663

is spin-forbidden because the lipids have a singlet ground state. However, this barrier is
overcome by the triplet oxygen attacking a fatty acid radical to form a peroxy radical.
This peroxy radical subsequently changes into a hydroperoxide by abstracting an (al-
lylic) hydrogen atom from another fatty acid molecule. This reaction is part of the chain
reaction and the rate-determining step:

The initial C-radical can form in several ways:


– by abstracting a hydrogen atom directly from the fatty acid (e. g., thermal). Only the
most weakly bonded hydrogen atoms can be used for this purpose, which must be
(at least) in an allyl position:

– by decomposing hydroperoxides into alkoxy, hydroxy, and peroxy radicals under


the influence of light, heat, or metal ions. The primary radicals abstract hydrogen
from a fatty acid and generate the C-radical:

The hydroperoxides are either already present in the natural material or created
from fatty acids by photooxidation or by catalysis from lipoxygenase (see below).

▶Figure 7.7 shows that the mesomerism of the intermediates gives rise to numerous
isomers; in all cases we obtain cis-trans and trans-trans-products from the natural all-
cis-compounds. As the temperature increases, the proportion of trans-compounds in-
creases. Radicals formed from a bis-allyl system such as the one of linoleic acid are ap-
proximately forty times more reactive than simple allyl radicals. They preferentially
form products, substituted at both ends of the bis-allyl system.
The oxidation products shown can also be observed with triple unsaturated fatty
acids. In these, however, the three double bonds do not act together, activating even
more strongly than two double bonds, but behave like two isolated bis-allyl systems.
As a result, terminal cis-double bonds remain in some products, which can form cyclic
664 � 7 Paint systems in art

Figure 7.7: Isomerization possibilities during autoxidative oil drying induced by mesomers of the allyl
radical.

peroxides. The radical can either form a hydroperoxide or further react to form bicyclic
compounds:
7.4 Oil paint � 665

Lipoxygenase reaction
Lipoxygenase is an enzyme that directly yields hydroperoxides from cis fatty acids.
There are several types of this enzyme, each of which targets specific positions, and
thus yields clearly defined hydroperoxides:

The resulting hydroperoxides can, by decay, provide the radicals necessary for the chain
reaction, as already shown for autoxidation.

Photo-oxidation
Through the photo-oxidation, we circumvent the spin barrier of direct oxidation men-
tioned before by using singlet oxygen as the reactant, which we can obtain from triplet
oxygen by irradiation:

Spin inversion can either be direct or via a sensitizer, which is activated in the first stage:
666 � 7 Paint systems in art

As a result of both variants, singlet oxygen can now readily attack double bonds of the
fatty acid in a cyclo-addition. In contrast to autoxidation, bond cleavage and linkage
occur in a concerted reaction:

We also obtain a complex mixture of isomers during photo-oxygenation since hydrogen


atoms on both sides of the double bond(s) can become the hydroperoxide-hydrogen,
▶Figure 7.8. Since no free radical occurs, the product mixture is composed differently
than in the case of autoxidation. From natural all-cis compounds, we obtain trans- or cis-
trans isomers. With increasing reaction time and temperature, the equilibrium shifts
to the side of the all-trans compounds. Symmetric hydroperoxides are not preferred.
Furthermore, six-membered cyclic peroxides can form in a concerted reaction:

The hydroperoxides formed in this process can also decay and provide radicals for the
chain reaction.

The chain reaction


At the core of oxidative oil-drying is the chain reaction of an alkyl radical with oxygen
forming a peroxy radical, which removes a hydrogen atom from another molecule of
fatty acid, thus regenerating an alkyl radical and forming a hydroperoxide:

In sum, one hydroperoxide forms from triplet oxygen per run.


7.4 Oil paint � 667

Figure 7.8: Isomerizations caused by photo-oxygenation during oil drying.

Compared to the alkoxy or hydroxyl radical, the peroxy radical is much less reactive
and abstracts only the most weakly bound hydrogen atoms of the reaction partners. The
binding energies of differently bound hydrogen atoms show that the target of the peroxy
radical must again be preferably an allyl or bis-allyl hydrogen. Since the resulting O−H
bond possesses an energy of about 376 kJ/mol, it becomes clear that saturated fatty acids
at room temperature cannot be attacked by the peroxy radical and that unsaturated
fatty acids are required for the drying process.

Decomposition of hydroperoxides, start of film buildup


We would not obtain a film if only isolated molecules of fatty acid hydroperoxides were
formed in oxidative drying. However, the hydroperoxides decompose relatively readily
into alkoxy radicals, which are available for polymerization reactions:
668 � 7 Paint systems in art

Heavy metals, heat, and light promote decomposition (▶Section 7.4.5) and increase the
reaction rate considerably. Peroxy radicals are formed due to the higher energy barrier
only when radical starters are added to the reaction mixture.
Two hydroperoxides can also react with each other to form peroxy radicals and
alkoxy radicals:

However, this reaction only occurs when hydroperoxide concentration has become suf-
ficiently large.

Polymerization of decomposed products


The first components of the resulting oil film are dimers, which form by recombination
of the various radicals present:

The recombination of two peroxy radicals leads to the unstable tetroxide I, which de-
composes under oxygen release into oxygen-containing fragments of various natures.
By abstraction of an (allylic) hydrogen atom from a hydrocarbon, C-radicals can emerge,
which provide further opportunities for recombination:

Adding a radical to an unsaturated hydrocarbon forms high molecular weight polymers,


retaining the radical character, and the addition continues as a chain reaction, ▶Fig-
ure 7.9, pathway I. Since radical attacks occur preferentially at allylic positions, typical
7.4 Oil paint � 669

Figure 7.9: Reaction pathways in the polymerization of oil films: radicals add to olefins and transfer the
radical character to dimers, oligomers, and others. Recombination with different radicals leads to oxy-
genated polymers.

linkage patterns with one to two oxygen atoms in the bridge emerge from all the above
reactions, ▶Figure 7.9, pathway II. Each intermediate radical can combine with hydroxyl
radicals, oxygen, or hydrogen donors to form various oxygen-containing products (al-
cohols, ketones).
The film-building polymerization is favored by a high content of unsaturated fatty
acids and high oxygen availability, as is the case during the drying of oil paintings. At
room temperature, peroxo bridges form preferentially, followed by ether bridges. How-
ever, C-C bridges only form in appreciable quantities at higher temperatures, at which
point the peroxides decompose, and further reaction possibilities come into play. We
exploit this in the hot polymerization process to obtain stand oil, ▶Section 7.4.4.

Decomposition of hydroperoxides, degradation and side reactions


The radical species formed during the decomposition of the hydroperoxides undergo
manifold reactions in addition to polymerization. These reactions are mostly undesir-
able, as they lead to low-molecular volatile decomposition products, which are often
aggressive, such as carboxylic acids, esters, and alcohols. In the food sector, we consider
these decomposition products to be distinctly unappetizing and associate terms such as
“rank,” “old,” and “fishy,” with them. As long as the oil film of paintings stays exposed to
atmospheric oxygen, hydroperoxides form slowly but steadily, and their decomposition
promotes further degradation of the oil film.
670 � 7 Paint systems in art

▶Figure 7.10 provides an overview of these reactions. The shown reactions begin in
the first phase of drying oils, the oxidative drying, while the content of hydroperoxides
and their radical decomposition products is high. However, they continue as long as a
painting remains exposed to atmospheric oxygen. Consequently, the already hardened
oil film is degraded by β-scission and Norrish scission [906].

Figure 7.10: Side reactions during the drying of oils: formation of various oxygen derivatives from the fatty
acids (possibly with cleavage of chains) and disproportionation of hydroperoxide radicals.

β-scission is a mechanism that cleaves bonds adjacent to a radical:


7.4 Oil paint � 671

Norrish scission attacks carbonyl compounds. Peroxy bridges of the film are homolyt-
ically cleaved by irradiation, yielding alkoxy radicals, which react further to form ke-
tones, alcohols, aldehydes, and carboxylic acids. Ether bridges and C-C-bridges are ox-
idatively cleaved and yield esters, alcohols, and ketones.
The resulting radicals can combine with hydroxyl radicals or oxygen to form var-
ious oxygen derivatives. The transfer of the radical function according to the vinylogy
principle leads to other products:

▶Figure 7.10 also shows the origin of the yellowing of oil films: Several conjugated dou-
ble bonds in α-position to a carbonyl group have a slightly yellowish color. Carbonyl
compounds form by recombining two peroxo radicals via an unstable tetroxide or by β-
and Norrish scission.
Besides the desired polymerization and the formation of low molecular weight
cleavage products under the breakage of the double bond, more side reactions such as
radical cyclization and radical epoxidation consume double bonds, ▶Figure 7.11. The
cyclization leads, under dimerization, to cyclohexene. Epoxidation often follows a radi-
cal attack on a double bond. At higher temperatures, as is common in boiling stand oil,
cyclizations do not occur radically but as [4+2]cycloaddition or Diels-Alder cyclization,
▶Section 7.4.4.

Hydrolysis of glycerides, pigment soaps


The reactions do not yet entirely stop after the oil film has thoroughly dried, which may
take several months. Over the following years and decades, the glycerol esters are hy-
drolytically cleaved, releasing glycerol and the (cross-linked) fatty acids [905]. Some fig-
ures give an impression: after 2 years, 20 % of the glycerides were hydrolyzed, and after
200 years, about 80 %.
Hydrolysis is vital for film quality since cross-linking in the film decreases, and the
resulting low-molecular substances either leave the film or act aggressively as free acids.
The hydrolysis is promoted by traces of moisture and basic metal oxides, as they occur in
pigments such as lead white. A well-known phenomenon is the increasing transparency
of paint layers in which lead white has been converted into lead carboxylates (“lead
soaps”). As organic lead salts, these are no longer semiconductors and are no longer
opaque white but transparent. Copper ions can also create copper salts with the released
carboxylic acids, including the formerly known, deliberately glazed copper green colors,
▶p. 225.
672 � 7 Paint systems in art

Figure 7.11: Other side reactions caused by a radical attack on double bonds, leading to their decrease.

7.4.4 Stand oils

Thickened or bodied oil is a viscous linseed oil type that we can prepare by heating
linseed oil in the absence of air (in a vacuum or a nitrogen atmosphere) to 250–300 ℃
and keeping this temperature for several hours, ▶Section [Link] at p. 659. During this
process, nonoxidative polymerization leads to a thick oil containing numerous cyclic
compounds and fatty acid oligomers [103, 189, 894, 907–913, 928]. In the reactions, about
2
3
of the double bonds disappear. In contrast to blown oils, a residue remains for regular
oxidative drying after application. Large portions of the naturally present antioxidants
are consumed during heating and cannot delay drying later. Due to pre-polymerization,
stand oils dry more slowly than regular oils and absorb less oxygen. This aspect is ad-
vantageous because the increase in the volume of the oil film during curing is not as
substantial, and the tendency to crinkle is reduced.
The product spectrum of the stand oil is diverse and not yet fully understood. Apart
from the great heat, a significant difference from oxidative polymerization is the almost
complete absence of oxygen, and thus, hydroperoxides. The initially present hydroper-
oxides are rapidly cleaved, and the peroxo-bridged dimers already formed also decom-
pose to lower oxygen dimers (R−O−R → R−O−R, R−R). The reaction is dominated by
(allyl) radicals formed by thermal dissociation:
7.4 Oil paint � 673

As a result of the radical formation and the possible mesomeric transfer of the radical
function into the neighborhood, a shift of isolated double bonds to a conjugated position
can occur. This modification changes the initial relative amounts of fatty acids, which
can complicate the analysis and identification of drying oils.
The product mixture is not, as in cold film formation with air participation, deter-
mined by oxygen bonds but by carbon bonds formed from the C-radicals: Radicals can
recombine or add to double bonds. The addition preserves the radical character and
leads to oligomers.

Polymerization via radical reactions


The carbon radicals II or III, mostly allyl radicals, formed at high temperatures do not
compose hydroperoxides but recombine or polymerize. The recombination of two rad-
icals leads to a carbon-linked dimer:

When the carbon radical adds to a double bond, a new radical forms, which maintains
a chain reaction that leads to polymers:

Polymerization via Diels-Alder reactions


In addition to radical reactions, a Diels–Alder reaction can also lead to polymerization by
cross-linking two unsaturated fatty acids to form a dimer, ▶Figure 7.12. If the oil contains
polyunsaturated fatty acids, further Diels–Alder reactions can form oligomers with the
remaining double bonds. Similarly, the double bond of the resulting cyclo-hexene can
undergo another Diels–Alder reaction, resulting in trimers and oligomers. In practice,
however, at most trimers are formed.
674 � 7 Paint systems in art

Figure 7.12: Example of a Diels–Alder reaction during stand oil boiling, leading to dimers. Possibly ad-
ditional Diels–Alder reactions lead to oligomers. Fatty acids without conjugated double bonds such as
linolenic acid must first thermally isomerize before they can function as dienes. The Diels–Alder product
still has free double bonds and can act as a dienophile for further Diels–Alder reactions.

The diene component of the Diels–Alder reaction requires conjugated trans-oriented


double bonds. Therefore, the unsaturated acids of linseed oil cannot enter the reaction
until isolated double bonds as in structure I have been isomerized to the conjugated
system IV. Isomerization can occur as shown above through thermally formed radicals
II and rearrangement to form III and IV, including also cis-trans isomerizations. The
reaction of II to III is favored because III is stabilized concerning mesomerism by the
conjugated double bond.
The conversion from isolated to conjugated double bonds is slow and rate-deter-
mining in the production of stand oil. However, if the oil contains conjugated fatty acids
from the outset, stand oils can be produced quickly and easily: Oils with conjugated
7.4 Oil paint � 675

double bonds can be produced at as low as 250 ℃, whereas oils with isolated double
bonds require about 300 ℃ [189].

Other low molecular weight products


Secondary heat treatment products are monocyclic and bicyclic fatty acids, which form
by intramolecular ring formation. The following compounds may give an impression;
the amount and position of double bonds and bridges are variable:

Polyunsaturated fatty acids can also form cyclohexadiene during heating. As diene com-
ponents, they lead to oligomers in a Diels–Alder reaction:

7.4.5 Effect of heavy metals, siccatives

Salts of heavy metals such as iron, copper, cobalt, or lead have such a marked accelerat-
ing effect on drying in oil painting that they are now commercially available as so-called
siccatives. Early oil painters were aware of this phenomenon and used paints containing
these metals as part of the pigment, first and foremost lead white. Since Old Master glaze
painters applied it in the underpainting and all other stages of the painting process, a
drying accelerator was thus evenly distributed throughout the work. Later, cobalt- or
manganese-containing paints took over this role.
The effect of the metal salts is based on the decomposition of the hydroperoxides
present to alkoxy and peroxy radicals, which initiate the chain reaction of film forma-
tion:
676 � 7 Paint systems in art

The metal cation is regenerated in the second step and acts as a catalyst. We need a metal
for this reaction, which possesses two oxidation states. In the first step, the reduced form
is effective. The theoretically possible direct oxidation of the oleic acids by the oxidized
form of the metal without the participation of hydroperoxides is very slow and does not
occur during oil drying:

7.4.6 Linseed oil varnish

Related to fatty oils and siccatives are products referred to as linseed oil varnish. Unfortu-
nately, the term has no strict definition. It partly refers to a series of more or less defined
linseed oil products, partly to the result of oxidative drying of a linseed oil product, i. e.,
the film. The linseed oil varnish is offered for various applications, such as protective
coating for garden or wood furniture, coatings, and as a base for commercial paints and
printing inks. In every case, it is more intended for coatings and commercial paints than
for artists’ paints.
Linseed oil varnish refers mainly to boiled linseed oil to which siccatives have been
added for faster drying. There are also “linseed oil varnishes” on the market, which
consist of (nonboiled) “linseed oil” mixed with turpentine and drying agents or “quick-
drying linseed oil.” The common understanding seems to be that of a boiled or nonboiled
linseed oil enriched with drying agents.

7.4.7 Technical improvement of colorants and painting agents in the nineteenth


century, paint tubes

In connection with oil painting, a product was developed that we encounter everywhere
today: the resealable metal tube for adhesives, foods, and countless other items. The
improvement originated in 1841 when it took the form of a tin tube and replaced the
(pig) bladders used until then for storing paint [604].
From the nineteenth century onward, it became usual for most artists not to pro-
duce their oil paints by themselves, as this had been common the centuries before. In-
stead, a new trade was established by paint dealers and paint manufacturers from the
nascent chemical industry or who worked as artists. They produced, distributed, and
7.4 Oil paint � 677

sold all kinds of pigments and prepared them as ready-to-use paint. The first begin-
nings of these ready-to-use paints existed already from the seventeenth century, but at
all times, the artists had the problem of storing the paint, wherever it came from, in a
suitable container if they did not want to make or buy a small amount of paint for each
session.
The requirements consisted of an airtight seal to prevent premature curing due to
oxidation and the delivery of small partial quantities. Until the early nineteenth cen-
tury, bladders were the principal containers, tightly sealed at their opening. To retrieve
the paint, they were punctured, the paint squeezed out, and the hole was closed with a
needle or a button. There were some disadvantages: some colors could decompose or
change, oxidation occurred after the first opening, smearing color around the hole was
the typical case, and bursting of the bladder when squeezing out the paint could never
be ruled out. Professional painters did not question these disadvantages. The need for
alternatives increased only in the nineteenth century when more and more amateurs
were engaged in painting and expected easy handling of the painting material.
In 1804, the English paint maker J. Rawlinson suggested fixing a cylinder in the open-
ing of the bladder and retrieving the paint through it. In 1822, J. Harris, an art teacher
from Plymouth, presented a tinned syringe that allowed the easy and clean squeezing
out of the paint through its nozzle. The experts recognized the benefits, but the devices
were too expensive. In 1840, W. Winsor, a painter and cofounder of Winsor&Newton,
a company still significant today, patented glass syringes. The colors were easily dis-
cernible through the glass, and the glass cylinders could be sealed by lacquered or im-
pregnated membranes for long-term storage. For use, the membranes were punctured
and fitted with nozzles and stamps. Winsor&Newton announced this container in 1840,
which was well received but also aroused competition. [605] displays some of the con-
tainers.
As a result, in 1841, T. Brown sold the solution with Brown’s patent collapsible color
tubes, which is still the standard for storing oil paints: the metal tube made of tin. How-
ever, the actual inventor is J. G. Rand, an American painter, craftsman, and inventor who
came to London where he punched the tubes from tin sheets and cold-formed the tin
discs into a tube by sudden pressure. London paint merchants and makers immediately
recognized these tubes as the definitive solution but did not like to see any other name
than their own on their paint tubes. According to [604], e. g., Winsor&Newton worked
on their tubes and agreed only after a public advertisement dispute with Rand that he
would supply the tin tubes for the company. Shortly afterward, he also became a sup-
plier for other paint manufacturers.
The malleable metal tube was quickly used to store all sorts of goods and was made
from other soft metals such as lead, but its origin was the search for a suitable container
for oil paints. Therefore, it was made of tin since tin, unlike other soft metals such as
lead, does not react with pigments, a fact that Rand knew as a painter. Reports of early
lead tubes in painting should therefore be regarded with care because they often refer
to other quickly found applications.
678 � 7 Paint systems in art

Today, tubes for colors are made of aluminum and are protected on the inside by
stoving varnish. Since aluminum could not be produced purely until 1825, and due to
the complex manufacturing process in the early nineteenth century, this material was
as expensive as gold. Therefore, it was not yet significant for the early tubes. Only in the
second half of the nineteenth century did the price of aluminum fall.

Metal tubes as a prerequisite for plein-air painting?


Ease of handling was not the primary consideration in the invention of the paint tube,
but rather its usefulness as a tightly closing container for color preservation. The in-
vention of the collapsible metal tube is sometimes credited for giving an impetus to the
open-air painting of the Impressionists (from the 1870s onward). However, according
to [604] and [606], contemporary chemical developments around oil painting may also
be significant. The commercially offered tube paints of this time exhibited considerably
improved drying behavior, mainly by adding manganese and cobalt compounds as sicca-
tives. New pigments with pure colors such as cobalt blue, emerald green, viridian, and
chrome yellow were bound to inspire the work of landscape painters. As binder analyses
show, Impressionist artists used the available, improved painting media. These include
prepolymerized stand oils added to purchasable paints and the commercial availability
of fast-drying, paste-texturing painting media (resin-oil mixtures), which were in de-
mand for impasto effects in the depiction of foliage and landscapes. Finally, it must be
taken into account that from the nineteenth century onward, many other aids such as
canvases, primers, portable easels, and brushes were buyable and made it increasingly
easier for the plein-air painter.
It fits into this more differentiated picture that, at first, portrait painters, rather than
plein-air artists used tube paints. In 1840, Winsor&Newton advertised the cleanliness
and low odor of the tube paints and not their suitability for outdoor use. In France, tube
paints were produced by Lefranc only from 1859 on, and in 1854, E. Delacroix still bought
paints in bladders. The taking of bladder paints was standard among French painters.
Moreover, 5 to 10 centimes were charged more for the expensive tubes, with an average
paint price of 25 centimes.
Furthermore, the idea of plein-air painting was by no means new; it had already
formed the basis of the painterly activities of the Barbizon School in the 1820s to 1850s,
as paintings by Corot and Rousseau show. The real impetus came from the age of enlight-
enment, which induced British artists to seek pictorial truths on natural locations and
produce oil sketches there, namely by R. Wilson, T. Jones, J. W. M. Turner, J. Constable,
or R. P. Bonington. In France in the early eighteenth century, representatives were
A.-F. Desportes, followed by P.-H. de Valenciennes, C.-J. Vernet, J. Coignet, A. Enfantin,
J. B. C. Corot, finally the Barbizon painters and the Impressionists. Callen [607] guides
very nicely through the history of plein-air painting. These landscape painters had
always used colors in bladders even when working outdoors and had overcome diffi-
7.4 Oil paint � 679

culties with the available material with their skills. Undeniably, tube paints decisively
simplified the outdoor work for nineteenth century artists. Today it is impossible to
imagine the artist’s workshop without tube paints.

7.4.8 Resins, resin balsam, turpentine oil

Apart from fatty oils and pigments, most of the materials we use for oil painting belong
to the large group of terpenes or isoprenoids. Solvents such as turpentine oil, varnish
resins such as dammar and mastic, and paint additives such as Venetian turpentine are
based on isoprene.

With a repeated linkage of C5 -isoprene units, according to ▶Figure 7.13, nature builds
bodies up to C45 . These parent bodies are subject to reaction possibilities such as cycliza-
tion, methyl shift, and hydroxylation, which yield a plethora of compounds. The low
molecular weight C5 representatives are volatile, highly aromatic compounds. Many
flavors and fragrance compounds belong to them and span a wide range, from fresh
lemons to fennel, camphor, cinnamon, or conifer aroma. The next higher C15 compounds
are hardly volatile and constitute aroma and bitter compounds, e. g., in grapefruit. As
molecular weight increases (C20 and above), we arrive at the resin acids. They serve
plants as wound closure, insecticide, or saponin, and we will look at them in more
detail in a moment. Even higher oligomers from C40 are important as yellow and red
carotenes, which are plant colorants in foliage, fruits, and vegetables, ▶Section 4.3.
While plants synthesize all levels, humans primarily produce triterpenes (steroid hor-
mones and bile acids).
C10 - to C30 -compounds are of interest as painting materials. They are produced by
many plant families and occur mainly in two forms:
– Resin balsams are viscous to solid masses deposited from living trees and consist
of di and triterpenes (resin acids, C20−30 ). The resin acids are dissolved in mono or
sesquiterpenes (C10−15 ).
– Deposited or fossil resins consist of partially polymerized resin acids and terpenes.
Amber is a beautiful example in the truest sense. Carpenters value the fossil resins
as components of valuable furniture varnishes.

The resin balsams provide the painter with important solvents and varnish resins [185,
Volume 1], [103]. Specific trees are cut to obtain them, and the sap that emerges is col-
lected [949]. The sap consists of two fractions:
680 � 7 Paint systems in art

Figure 7.13: Biosynthesis of isoprenoids from isopentenyl pyrophosphate (IPP) and dimethylallyl pyrophos-
phate (DMAPP) [274, p. 89], [273, 275]. The paths relevant to painting that lead to solvents (turpentine oil),
resins, and carotenoid colorants are indicated.

– Liquid monoterpenes. The resin balsam of conifers (pines, spruces, firs, larches)
provides turpentine oil.
– Solid di and triterpenes (resin acids, balsamic resin). We can derive rosin from
conifer balsam, the dammar resin from certain Indonesian tree resins, and the mas-
tic from pistacia species.

The resin is initially liquid when it flows out but rapidly thickens as soon as the most
volatile monoterpenes have evaporated. The nonvolatile balsamic resins remain behind
7.4 Oil paint � 681

and partially crystallize so that the sap becomes cloudy. We can obtain the two main
components in pure form by distillation and subsequent purification. Since the distil-
lation technique has been known in Europe since the sixteenth century, oil painting
developed further only with the availability of this solvent.

[Link] Turpentine oil


Turpentine oil represents the purified, volatile distillate from the crude resin of conifers.
Pine species are used almost exclusively, as these trees proliferate, are widely dis-
tributed, and produce high resin yield (up to about 3 kg per year), lowering the turpen-
tine oil cost.
The composition of turpentine oil varies greatly with its origin, usually consisting
of monoterpenes of approximate relative abundance [185, Volume 1], [103, 963, 965]:

α-pinene > β-pinene > Δ3 -carene > limonene > myrcene > β-phellandrene

The amount of α-pinene can be as high as 70 %. Due to its structure, turpentine oil is
nonpolar but more polar than pure hydrocarbons due to its oxygen functions. Therefore,
we can use it as an efficient solvent for fatty oils (linseed oil) and natural resins (rosin,
dammar, mastic). It is also applicable for dispersing waxes.
High-grade turpentine oil is distilled several times and evaporates without residue.
The double bonds can give rise to oxidation and polymerization reactions (resinifica-
tion) during storage but without a regular film being formed.

[Link] Rosin, dammar, mastic


The solid components remaining after the distillation of the volatile resin oils from the
resin are the actual balsamic resins. As a result, we obtain several well-known products
based on the resin’s origin, ▶Table 7.7. Balsamic resins consist of a complex mixture of
682 � 7 Paint systems in art

Table 7.7: Botanical origin of important balsam resins [103, 963].

Type Product

Coniferae (Pinaceae)
Pinus (pine) Common or Bordeaux turpentine, rosin
Abies (fir) Strasbourg turpentine, Canada balsam
Larix (larch) Venetian turpentine
Leguminosae
Hymenaea, Copaifera Copal resins, copaiba balsam
Dipterocarpaceae
Hopea, Shorea (winged fruit plants) Dammar resin
Anacardiaceae
Pistacia (pistachio) Mastic resin

resin acids with di, tri, or tetraterpene skeletons and a proportion of fatty acids and fatty
alcohols [963, 965].

Pinaceae resins
They are best known to us in terms of their composition. Although it is subject to varia-
tion, we can indicate the following rough proportions:

pinus species:
Levopimaric acid/pallustric acid > iso-pimaric acid > (neo-)abietic acid > dehydroabietic acid, pi-
maric acid, sandaracopimaric acid

abies species:
Abienol > abietic acid ≥ levopimaric acid/pallustric acid > neo-abietic acid

larix species:
Larixyl acetate > iso-pimaric acid > levopimaric acid/pallustric acid > (neo-)abietic acid ≥ epi-
manool/larixol

During distillation, isomerization of the double bonds occurs so that the original com-
position changes, and we, e. g., no longer find levopimaric acid in the rosin but more
abietic acid.
Carboxylic acids with abietane and pimarane skeletons dominate the pinus resins.
They do not show a high tendency to polymerize and form soft resins.
7.4 Oil paint � 683

Abies and larix balsamic resins also contain compounds with the bicyclic labdane skele-
ton:

The labdane skeleton of abies species can polymerize and lead to the hardening of the
resin over time. Polymerization occurs via the diene structure [964], primarily as a 3,4-
addition and to a minor extent as a 1,4-addition, since the internal double bond is steri-
cally shielded:

Leguminosae resins
They form the basis of copal resins and are also based on the labdane skeleton. In addi-
tion, there are soft, nonpolymeric resins (copaiba balsam) and hard resins (copal resins)
based on labdadiene in which polymerization has taken place through the conjugated
double bonds. Remarkably, these resins contain almost only the enantiomers of the con-
stituents of the pinaceae species. However, since many of the resins were studied before
the era of GC/MS, our knowledge is less precise than that of conifer resins:
684 � 7 Paint systems in art

Dammar resins
They are built up from triterpenes. A nonpolar, alcohol-insoluble fraction (β-resene)
consists of poly-cadinen, and the remainder (the polar, alcohol-soluble, and partially
acidic fraction, α-resene) contains hydroxy- and oxo-derivatives of tetracyclic and pen-
tacyclic triterpenes with the skeletons of dammarane, oleanane, ursane, and hopane
[103, 948, 949, 951, 963]:
7.4 Oil paint � 685

Dammar resin is sparingly soluble in water and other solvents but dissolves in turpen-
tine oil and melts at temperatures above 80 ℃. It has been a classic varnish resin since
the nineteenth century; in ▶Section 7.4.10, the role of varnish is delineated in more
detail.

Mastic resin
This resin was already known in antiquity. It was obtained from trees of the species
pistacia, native to the Mediterranean region, especially the island of Chios, and thus
occurs in the surroundings of the ancient cultures. Similar in composition to dammar
resin, it contains high proportions of compounds with the euphane and oleanane skele-
ton, as well as one polymer (cis-1,4-poly-β-myrcene) [103, 951, 955, 963]:

Mastic resin has a melting point of around 100 ℃ and is slightly soluble in water and
other solvents but readily in turpentine oil. This solution has been used as a resin var-
nish since the seventeenth century, ▶Section 7.4.10.
686 � 7 Paint systems in art

[Link] Aging, oxidation of resins


The protective effect of varnishes with dammar or mastic resin is controversial since
the easy oxidizability of the terpene ketones leads to numerous low molecular weight
degradation products over time [948, 950–953]. ▶Figure 7.14 shows two pathways that
could cleave the terpene ring systems. Both follow the typical chain reaction of radical
oxidations:

Figure 7.14: Typical reactions during oxidative degradation of dammar and mastic resin varnishes. The
figure shows a typical partial structure of triterpenes and possible subsequent reactions.

The initial radicals preferentially form adjacent to double bonds or carbonyl groups.
Recombination or the formation of new double bonds breaks the radical chain. As a re-
sult, alkenes, aldehydes, carboxylic acids, and carbonyl groups are created. In the mastic
resin, up to six oxygen atoms are introduced into a molecule in this way. Typical oxida-
tion products are
7.4 Oil paint � 687

Interestingly, oxidation occurs rapidly and independently of light exposure; only the re-
action rates vary. Thus, after 7 weeks in the light, similar amounts of degradation prod-
ucts were found as after 31 weeks in darkness.
As a result of oxidation, the skeletons of the triterpenes become reduced, modified,
and highly functionalized. Polymerization of the intermediates hardly takes place due
to the rapid further reaction of the radicals to carbonyl groups. As a result, the oxy-
genated compounds are hydrophilic and often acidic, leading to the well-known solubil-
ity of aged varnishes in alcohols.
Significant for the conservation of old paintings are brittleness and yellowing of the
varnish caused by the decomposition of the resin components; altered or faded colors
are the consequences. Yellowing occurs especially when the varnished paintings are
stored in the dark; this can be caused by light-induced oxidation and bleaching of the
colored degradation products by sunlight. The coloration of the degradation products
is caused by more or less extended double bond systems in conjugation with carbonyl
groups, as they emerge during autoxidation, ▶Figure 7.15.
Many of the disadvantages mentioned should no longer occur in modern varnishes
based on cyclohexane or similar substances. However, their very short modern product
cycles do not yet allow any statements over a more extended period, ▶Section 7.4.10.

7.4.9 Other solvents: benzines, turpentine substitutes

Today, in addition to vegetable balsams such as turpentine oil, we can also buy solvents
based on mineral oil. These include “wound gasoline,” “wash gasoline,” nitro thinner,
and turpentine substitutes. These terms refer to liquid mixtures of aliphatic and aro-
matic hydrocarbons that result from gasoline production in petroleum refineries. The
crude gasoline fraction (naphtha, boiling point 30–220 ℃) is further separated by frac-
tional distillation, giving us different groups of mineral spirits [914].

Aromate-free aliphatic solvents


Aromate-free aliphatic solvents consist of lower aliphatic compounds and have low boil-
ing points and flashpoints. They are classified according to their boiling range and can
be obtained chiefly freed of aromatics by fractional distillation:
688 � 7 Paint systems in art

Figure 7.15: Emergence of colored, quinoid, or conjugated structures during oxidative degradation of resin
varnishes [956, 957].

– petroleum spirit 40–60 ℃: n-pentane, 2-methyl-pentane, cyclopentane, and cyclo-


hexane
– petroleum spirit 60–80 ℃: n-hexane
– petroleum benzine 100–140 ℃: iso-octane, cyclohexane to cyclooctane, n-hexane,
n-heptane
– petroleum spirit 140–165 ℃: n-octane to n-decane, iso-octane, iso-decane, cyclooc-
tane, cyclononane

Aliphatic solvents containing aromatics


These solvents are mixtures of aliphatics and aromatics:
– naphthabenzine 100–140 ℃: n-hexane to n-nonane, iso-heptane to iso-nonane, cy-
cloheptane, cyclooctane, toluene, xylenes
– crystal oil 135–180 ℃: n-octane to n-undecane, iso-octane to iso-undecane, cyclooc-
tane, cyclononane, xylenes, mesitylene, cumene, and other aromatics
– turpentine substitute, turpentine oil substitute, white spirit 140–200 ℃: n-octane
to n-dodecane, iso-nonane to iso-dodecane, cyclooctane to cyclodecane, tri- to pen-
tamethyl-benzenes

Today, they are also offered dearomatized by hydrogenating the aromatic components.
7.4 Oil paint � 689

Solvents rich in aromatics


They consist almost exclusively of aromatics, and thus have excellent dissolving power
but, unfortunately, also high toxicity:
– 160–182 ℃: tri and tetramethyl benzenes, cumene.

7.4.10 Varnish materials

Before we consider the types of varnishes used, we must keep in mind what function a
varnish should perform:
– A varnish gives the painting more surface gloss, evens out microscopic irregularities
in the painting layer, and provides a smoother surface. Consequently, the amount
of light is reduced, which would otherwise be diffusely reflected and scattered by
fine inhomogeneities, ▶Section 1.6.8. Instead, there is an increased depth of light
and directional reflection, which manifests as surface gloss.
– The varnish adds depth of color and color clarity to the image due to the depth light.
In addition, the reduced scattering causes more light to pass through the painting
layers and emerge as colored depth light, and the colors appear darker and richer,
▶Section 1.6.8.
– The varnish gives the image more sharpness of detail because the varnish encloses
all pigment bodies in a transparent body with a uniform refractive index. The inci-
dent light is thus refracted uniformly into the varnish and not to variable degrees
at different pigment grains.
– Layers of varnish protect the painting from environmental influences.

[958] illuminates these points in more detail and discusses the advantages and disad-
vantages of varnishes.
Few information sources about the early varnishes exist. However, known and
available materials always served as a basis. ▶Figure 7.16 indicates essential key
data [945]. We know that early varnishes consisted of oils thickened or fused with
various resins to form thick layers with a high tendency to yellowing and browning.
Then, in the seventeenth century, the first spirit-based varnish appeared, which could be
applicable from today’s point of view: mastic dissolved in turpentine oil, which became
the crucial painting varnish in the nineteenth century and early twentieth century [959].
Mastic also shows a slow yellowing. From 1827, dammar resin played a role, but it was
little appreciated in the English and French regions. It was not established until the late
twentieth century, especially in German-speaking countries. It yellows less than mastic.
Both resins are initially soluble in turpentine oil, the solubility progressively decreases
with aging, and the solubility in alcohol increases.
690 � 7 Paint systems in art

Figure 7.16: The development phases of varnishes [945].

Natural resin varnishes


The most critical application of dissolved balsamic resins in painting is as a varnish,
which came into practice in the seventeenth century. At that time, turpentine oil distil-
lates became available, needed to dissolve the resins.
The solvent evaporates from the resin solutions a short time after application, and
the remaining balsamic resins spread thinly on the painting, forming a first solid, pro-
tective film. The resin film initially dries up purely physically, and cross-linking hardly
occurs in the further course [948, 950–953]. The actual oxidative-chemical drying of
the fatty oils takes place in and under the resin film in the course of the following
weeks.
We have already learned about the chemistry of dammar and mastic resins in detail
in ▶Section 7.4.8 at p. 679 connected with balsamic resins and touched upon the ques-
tion of the aging of resin films. Numerous investigations [948, 950–953] show that the
protective function of a natural resin film diminishes after a time and finally ceases to
exist. A detailed investigation into the general suitability of dammar resin as varnish is
given in [949].

Synthetic polymer varnishes


Synthetic polymers are repeatedly mentioned as an alternative to the natural resin var-
nishes with their recognized disadvantages. In Koller and Baumer [946], some high poly-
mer materials of the twentieth century (acrylates, polyesters) are discussed, but not as
an alternative to the conventional resin varnishes.
In this way, e. g., the low-molecular natural resins dry without tension because the
small molecules can arrange themselves freely as the solvent evaporates. On the other
hand, polymers are hindered in their mobility and remain in a “frozen” state of tension
after evaporation, so the addition of plasticizers is necessary. In the case of polyacrylate
polymers, copolymers can be used instead.
7.4 Oil paint � 691

Cyclohexanone varnishes (ketone resins)


Cyclohexanone varnishes belong to low-molecular materials, but they are also nega-
tively evaluated in [947]. Their mechanical and film-forming properties, redissolubility,
and tendency to embrittlement also do not meet the requirements of a varnish. Never-
theless, they are valued for the restauration of Old Master’s resin varnishes [954].
The chemical structure of these varnishes, which have found a certain distribution
since 1952, is not fully known; their resin basis forms two types of oligomers with the
approximate structures shown in ▶Figure 7.17 [947, 954]. Both types are built by con-
tinued condensation from cyclohexanone or methyl-cyclohexanone, splitting off wa-
ter molecules and forming single or double bonds. One molecule incorporates up to 12
monomers.

Figure 7.17: The two types of cyclohexanone resins and the synthesis of type II resins [947, 954].

7.4.11 Pigment degradations

We have already discussed the undesirable change in pigments affecting the color over
long periods based on damage to mural paintings, ▶Section 7.3.5. Oil paintings, draw-
ings, and watercolors are stored in a more protected manner. Therefore, the damage is
often not so pronounced. Nevertheless, there are several undesirable changes here as
well, of which we can only look at typical representatives, such as
– the (total) loss of yellow lakes (visible in areas of mixed green leaving blue trees and
plants)
– the (total) loss of red lakes (visible in pale faces and dull robe fabrics)
692 � 7 Paint systems in art

– the (total) loss of blue color from Prussian blue or smalt (visible in silvery-gray skies
and changed tonality of the painting)
– the aging of the linseed oil medium (visible in the vanishing of the painting subject
into hard-to-read, formless semidarkness)

Color changes are caused by pigment or binder degradations, by the reaction of pig-
ments with each other or with binders. In the following, we will look at examples for
which a basic level of knowledge is already available. Overall, however, pigment and
paint damages are a current research topic in art technology; see, e. g., [389, 390].

[Link] Reactions of iron pigments


While iron oxide pigments are very stable, the iron blue pigment (Prussian blue) is a
different case. Shortly after introducing this intrinsically valuable blue pigment at the
beginning of the eighteenth century, the tendency of the pigment to fade became appar-
ent. Among painters, this phenomenon was perceived. Why they used the pigment re-
gardless has been outlined in ▶Section 1.3.4. Consequently, we found examples of this in
the works of R. Wilson, A. Watteau, Canaletto around 1730, T. Gainsborough around 1750,
or Tiepolo around 1750, notable, especially in bleached-out parts of the sky, which are
nowadays too light, turning greenish or whitish. Börner, Kirby, and Saunders [431–433]
confirm the tendency of Prussian blue to fade due to photochemical reactions.
Systematic investigations showed that fading only occurs with white mixtures, not
pure pigment. Unfortunately, the pigment can hardly be used in isolation due to its high
coloring power. Partly responsible for the poor durability of the early pigment were
obscure manufacturing processes and their parameters. They were determined by trial
and error, their influence on the pigment quality was unclear, and the addition of fillers
was not traceable.
The modern pigment is listed as moderate to durable, except in very thin glazes
or mixtures with a high content of white. Good durability is the result of a controlled
production from pure starting materials.

[Link] Reactions of cobalt pigments


Smalt, a cheap cobalt blue glass, was used to supplement the limited blue range for a long
time. Unfortunately, while cobalt blue pigments are very stable, the same can not be said
of smalt. Consequently, it soon became apparent that the glass was prone to bleaching.
The result is visible, e. g., in paintings by P. Veronese, B. Bruegel, or J. Beuckelaer, all from
around 1560–1570, or T. Gainsborough around 1750 [426, 552, 559]. Optical damage in-
cludes total vanishing of blue sections, graying and discoloring of delicate bluish-purple
shadows, and yellow-brown discoloring of the medium. When used in skies and to tint
shadow areas, these changes often alter the entire tonality of the painting.
According to [426–428], color bleaching occurs due to the structural loss of the potas-
sium glass, the fundamental matrix of smalt. In smalt, cobalt is present as a tetrahedrally
7.4 Oil paint � 693

coordinated CoII cation whose LF transitions are symmetry-allowed and intense due to
tetrahedral symmetry (CoO4 chromophore). However, as soft glass, potassium glass is
notoriously susceptible to moisture and acids, which slowly dissolve out the alkali and
alkaline earth cations, especially potassium in the case of smalt. Potassium stabilizes
the tetrahedral structure so that after the migration of K, the chemical environment of
CoII changes, and the coordination number increases up to the octahedral coordination.
However, this means that a symmetry-forbidden LF transition is now responsible for the
color, which is considerably weaker:

Alterations in the medium’s color are accompanied by complex changes triggered by the
formation of potassium soaps.

[Link] Mercury pigment reactions


Blackening of vermilion is a well-known phenomenon [395]. Red vermilion can be sta-
ble for hundreds of years, while in other paintings, it can form a silver-gray or black
crust that severely disfigures the painting’s content. Examples span all centuries, from
G. del Ponte around 1420, P. Uccello around 1440, B. Gozzoli around 1460 to J. Jordaens
around 1620, or A. Cuyp around 1650.
Investigations showed that impurities from the wet production favor the black-
ening; nevertheless, the more stable vermilion from the dry production also shows
blackening, sometimes next to well-preserved parts in the same painting. In most cases,
black and white degradation products were found side-by-side. It has been shown that
the white decomposition product mercury(I) chloride (calomel) emerges in coopera-
tion with chloride from the air or dirt, light, and moisture. Earlier attempts formulated
reactions such as these, additionally yielding black mercury sulfide:

Newer experiments suggest that the reaction proceeds via kenhsuite and corderoite up
to elemental mercury [389, 396]:
694 � 7 Paint systems in art

In addition to the purity of the cinnabar, painting methods have a decisive influence on
whether it blackens or not. Examples show that cinnabar has been preserved where it
was protected by additional layers of paint, which either reduce the amount of light or
form a barrier against, e. g., chloride ions.

[Link] Reactions of cadmium pigments


The bright cadmium yellow pigments were frequently used in the nineteenth and twen-
tieth century by impressionists. Nowadays, in some important artworks, a discoloration
can be observed, e. g., in works by Van Gogh, Ensor, or Matisse [389]. The mechanism
behind is a photo-oxidation of cadmium sulfide to cadmium sulfate in the presence of
oxygen and moisture. Depending on the environmental conditions, manifold other cad-
mium compounds may be produced:

The reaction is not yet completely understood, some of the compounds found may be
residues or additives from the manufacturing process. Also, the influence of the crystal
structure and stoichiometry on stability is not yet clear.

[Link] Reactions of arsenic pigments


Realgar is a bright orange-red pigment, which in some paintings shows significant
change to a yellow powder, e. g., in J. Tintoretto’s Gonzaga cycle, whose expressive
yellow highlights on fabrics and in flames initially showed milder orange tones [549].
Wallert [408] contains a formulation for a transformation of realgar to yellow orpiment:
7.4 Oil paint � 695

However, according to [60], orpiment is thermodynamically more unstable, so the trans-


formation must be formulated with yellow para-realgar as the final product:

In this mineralogically well-known transformation, the weak As−As bonds in the realgar
break, and free arsenic forms, destabilizing the realgar structure and combining with
the surrounding arsenic and sulfur atoms to form the powdery structure of para-realgar.
Despite the same molecular formula, para-realgar has a different structure than realgar.
Arsenic (white arsenic(III) oxide), also mentioned in the sources, can be formed as an
oxidation product.

[Link] Reactions of lead pigments – Changes in binder


Burmester and Krekel [549] describe a change in a picture cycle by J. Tiepolo, affecting
the binder and effective over large areas: the vanishing of the picture subject into a
semidarkness where contours or colors are difficult to discern. This example is only one
of the countless other paintings of the most diverse painters and epochs, e. g., the Dutch
of the seventeenth century such as Rembrandt, A. van der Neer, J. van Goyen, J. Steen,
F. Hals, J. Vermeer, N. Berchem, or also V. van Gogh 1888. Affected are oil paintings from
the fifteenth century to the twentieth century; on canvas, paper, wood, or copper; even
layers of painting in egg tempera or varnish layers can show this type of damage. The
mentioned visual changes include physical damages to the painting material, which
may occur in different combinations: microscopic holes of 100–200 µm size distributed
over the painting, streaky dark spots, incrustations, and spalling of parts of the painting
layer. Facing many of these disturbing, damaged paintings, restorers, technicians, and
chemists attempted explanations already at the beginning of the twentieth century.
The chemist, A. P. Laurie, explained the increased transparency of whole layers of
paint with an age-related increase in the refractive index of the linseed oil medium.
Fresh linseed oil has a refractive index of about 1.48, which increases to about 1.6 with
aging (▶Table 1.17 on p. 53). A pigment’s opacity depends on the difference between its
refractive index and that of the medium, ▶Section 1.6.8. Since the refractive index of,
e. g., malachite, verdigris, bone black, and smalt is close to that of the fresh linseed oil
(▶Table 1.20 on p. 71), initially covering layers of these pigments can appear transparent
today even with a slight increase in the refractive index of the medium.

Lead and zinc soaps


This explanation is correct for the low refractive index pigments mentioned. However, it
does not explain why painting layers with high refractive pigments such as white lead,
696 � 7 Paint systems in art

whose refractive index even in aged linseed oil is still clearly above the medium, are
affected. At the beginning of the twentieth century, chemist, A. Eibner, argued that pig-
ments and siccatives release lead cations. These form low-refractive lead soaps (lead
salts of fatty acids) with free fatty acids of the linseed oil and thus cause transparency.
Subsequently, it was also observed that damages were caused by other lead compounds
such as lead-tin yellow or red lead. In the late twentieth century, the subject of “lead
soap” and, more generally, “metal soap” became the subject of broader research deal-
ing with the phenomena and damage patterns of aging of painting layers. The interim
results display the painting layer as a system full of dynamics [915–920]:
– The refractive index of fresh linseed oil of 1.48 increases to about 1.6 during aging,
making low refractive pigments transparent or translucent.
– Lead soaps consist mainly of stearic acid, palmitic acid, and azelaic acid lead salts.
They form amorphous voluminous aggregates, migrate to the surface and breach
it. This increase in volume leads to physical changes in or on painting layers, e. g.:
Peeling and flaking, lifting, micro-perforation, cratering, efflorescence, haze, and
incrustations.
If underpainting is causative, the damage may be local or picture-wide if the metal
soaps form in the primer.
– Lead soaps lead to significant visual changes: paint layers containing lead white, im-
primaturas, and primers become transparent, ▶Figure 7.18. Transparency emerges
by lowering the refractive index from 1.9–2.1 for lead white to 1.50–1.55 for the
formed lead soaps. This value is close to linseed oil and makes saponified lead white
practically transparent since scattered light is partly or entirely suppressed. The ac-
companying visual change depends on the color of the affected layer and the layer
structure as a whole:
– Underlying layers become discernible through shape and contour-giving upper
layers, primarily when low opaque pigments such as smalt are used in skies.
Underpainting and pentimenti shine through. The image becomes difficult to
discern due to loss of form and disappearance of the intended light-dark con-
trast.
– White grounds or light imprimaturas disappear optically and let the often dark
or brown-gray deeper layers shine through, or even the picture carrier itself,
e. g., dark irregular wood. Finally, the painting subject vanishes into the ground.
– Thick layers of lead white are less vulnerable than thin ones.

All effects work together to affect a complex pattern of damage. So far, mainly soaps with
the participation of lead or zinc cations have been identified, although it is difficult to
distinguish metal cations from a pigment or a siccative. For example, damage patterns
with zinc soaps can be found in images of the nineteenth century and twentieth century
when zinc white was intensively used. The source of lead cations is mainly lead white,
but also lead-tin yellow, red lead, or siccatives. Copper and potassium soaps occur to a
7.4 Oil paint � 697

Figure 7.18: Increase in the transparency of whole painting layers with lead soaps [915, 916, 919]. The
change becomes visible when no white painting layer exists under the layer that has become transparent.
The depth light that now appears makes the lower painting layers or even the entire primer visible. If the
primer has become saturated with oil, it also turns transparent. Besides, the often dark primer creates a
general semidarkness. In addition, the deeper painting layers no longer fit the motifs visible at the top,
resulting in a lack of contours and poor legibility. Superficial dirt and yellowing binders exacerbate the
problem. ▶Section 1.6.8 describes the consequences of scattering phenomena. (a): Freshly dried painting.
The primer in the example is calcium carbonate, n = 1.57 (white). Imprimatura and painting layer consist
of aged linseed oil, n = 1.60, lead white n = 1.9, yellow ocher n = 2.3, and a blue pigment. The color effect
is as intended by the painting layer and the imprimatura. (b): Aged painting. The primer has absorbed the
linseed oil. The difference in the refractive index has disappeared, and the primer has become practically
transparent. Lead white has been transformed into lead soap and is in linseed oil also transparent, n ≈
1.50. Light can penetrate unobstructed to the image carrier and strongly distort the visual impression as
dark depth light.

lesser extent. As fatty acids, mainly palmitic and stearic acids, were found, to a small
extent, also azelaic acid as an oxidative degradation product, ▶Figure 7.19.
The fatty acids are provided by hydrolysis of the triacylglycerols (TAG) of the oil
medium. Investigations on Dutch pictures of the seventeenth century show dark discol-
orations, which trace the grooves of the wooden substrate. The lime base has accumu-
lated thick layers in these grooves and has absorbed oil from the medium. In addition
to the darkening mentioned above, the primer has become transparent by matching the
refractive index to oil. It is assumed that such thick layers of a porous lime primer soak
up oil like a sponge. During the aging process, the necessary free fatty acids are formed
by hydrolysis at high concentrations at specific points and cause localized damage. This
example shows that the concrete damage patterns are strongly dependent on the struc-
ture of the painting layers and can thus be assigned to an underlying school, epoch, or
artist group.
698 � 7 Paint systems in art

Figure 7.19: Formation of transparent lead soaps from fatty acids and lead cations. The fatty acids orig-
inate from the hydrolysis of triacylglycerides of linseed oil or the oxidative cleavage of fatty acids (aze-
laic acid), the lead cations from lead-containing pigments, typically lead white, lead-tin yellow, or lead-
containing siccatives [915–917].

The exact processes are still the subject of research. Casadio et al. [915, Chapter 3] and
Hermans [920] summarize the results. According to them, in a first step, aged linseed oil
with soap-forming pigments represents an ionomer, a cross-linked anionic polymer with
carboxyl groups from hydrolysis and cleavage reactions. Metal cations released from the
pigment diffuse through the ionomer until dynamic electrostatic-chemical equilibrium
has been established, and the carboxyl groups are bound to metal cations I:
7.4 Oil paint � 699

The partners in these metal carboxylates are not permanently bound, and the metal
cations can diffuse and form new loose coordinative bonds.
In a second step, hydrolysis of the esters (triacylglycerides) contained in drying
oils occurs, either spontaneously by moisture or by the involvement of diffusing metal
cations. Cations such as PbII can act as Lewis acids and be coordinately bound by car-
bonyl groups of the esters. Due to their electron-withdrawing properties, the bound
metal cations weaken the ester bond, and thus facilitate hydrolysis by moisture. The
metal cations can bind to the resulting fatty acid carboxylates (FA) and form long-lasting
metastable systems.
In the third step, soaps can form rapidly when the equilibrium is disturbed, and a
more significant part of free fatty acid carboxylates becomes available, either by hy-
drolysis or by adding, e. g., beeswax or aluminum stearates. Saponification proceeds
rapidly and irreversibly since the metal soaps formed are insoluble in linseed oil (struc-
ture II, see above). After the formation of crystallization nuclei, free fatty acid carboxy-
lates and metal cations diffuse to the nuclei and enlarge them to macroscopic aggre-
gates. The exact processes depend on the concentrations of all the partners (such as
fatty acid carboxylates and metal cations) and the diffusion, hydrolysis, and crystalliza-
tion rates. Ultimately, the cleavage processes leading to the free fatty acids are complex
and depend on conditions in the painting layer and environmental parameters. For ex-
ample, high humidity leads to photolysis of fatty acids into shorter dicarboxylic acids
such as azelaic acid, while higher temperatures accelerate hydrolysis to saturated fatty
acids.

Other metal soaps


Especially in connection with smalt, potassium soaps could also be found, formed when
the glass deteriorates and K migrates into the medium where it reacts with fatty acids.
This process is assumed responsible for the brown discoloration of the medium and
superficial clouding. Finally, copper soaps have long been known as reaction products
of copper pigments with oil.

Further damage patterns caused by lead pigments


In addition to the damage patterns mentioned above, others include the formation
of delicate white veils or dense white crusts of complex salt mixtures when smalt is
present. The salts have been identified as lead potassium sulfate and calcium oxalate.
Cations of lead and potassium are derived from smalt potassium glass, PbII also from
lead white, the necessary sulfur from the atmosphere. The formation of the lead soaps
promotes this reaction, as it facilitates the dissolution of lead white and promotes mo-
bilization of lead cations:
700 � 7 Paint systems in art

Red lead can bleach by the reaction of photo-excited Pb3 O4 that reduces PbIV to PbII , in-
duce decarboxylation of the oil binder, and form white lead carbonate from the released
CO2 [389]:

[Link] Reactions of red and yellow lake pigments


Frequent and conspicuous examples of the degradation of organic pigments appear
in the area of yellow and red lakes [401–403, 549, 559]. The yellow lakes include laked
flavonoids such as Stil de grain from rhamnus berries or yellow resins such as gamboge.
We have already addressed the importance of yellow lakes for green mixtures and their
problems, ▶Section 1.3.5. Typical, distinct visible changes occur in paintings over time
caused by the problems mentioned there.
Eye-catching is part of painted foliage or whole forests and landscapes, as well as
fruit, which in extreme cases appear gray-blue or streaky-brown today, as in paintings by
J. v. Huysum, J. Steen, M. Hobbema, C. Lorrain, P. Lastman 1618, A. Cuyp 1650, J.-B. Geuze
at the end of the eighteenth century. Here are the yellow lakes of the green mixture
wholly decomposed. Yellow lakes lose color and leave behind white material from the
substrate and unknown decomposition products, making the painting appear blanched,
chalky, or cloudy.
Fortunately, the changes mentioned above are not ubiquitous. The majority of Old
Master paintings do not show signs of degradation at all or only to a moderate extent,
without noticeably impairing the effect of the painting. It is, therefore, a goal of today’s
research to clarify the exact circumstances under which damage occurs. Stable yellow
pigments are lead-tin yellow (opaque), Naples yellow (opaque), yellow ocher, and in
more modern times, chrome yellow.
The red lakes include madder lake, crimson, and lac dye. Effects of fading of red
lakes are detectable in paintings from the fourteenth century on, e. g., by J. di Cione 1370,
L. di Monaco 1414 (lac dye), B. Daddi, A. Gaddi, and other Florentine and early Italian
paintings of the fourteenth century and later the fifteenth century, by Tintoretto 1580,
Campana 1590, J. Reynolds 1760 (cochineal lake), or T. Gainsborough 1770 (cochineal
7.4 Oil paint � 701

lake). While yellow lakes are often used in landscape and nature depictions, deteriorated
red lakes result in pale white faces, lacking freshness, liveliness, modeling, and dull-
colored clothing and draperies, which in the circles of the noble patrons were hardly
worn like this.
In the case of red lakes, influencing variables and their effects have already been
systematically investigated [401]. However, the question of the substrate or the metal
used for laking, in general, is difficult to answer. Before the nineteenth century, the sub-
strate of lake pigments was white and translucent in the binder, usually aluminum ox-
ide hydrate Al2 O3 , AlOOH, Al(OH)3 , or Al2 O3 ⋅ n H2 O. Often a Ca salt such as CaCO3 was
added instead of or in addition to the precipitating reagent, forming a Ca Al salt. Excess
of the Ca salt would reach the final product as filler material. Additions of clay or other
white inert material were also common, so the final aluminum content could have been
very low, possibly resulting in a correspondingly low adhesion of the colorant to the
substrate. On top of that, Ca instead of Al in the complex influences the lake’s achieved
color. Various influencing factors were identified and their effects investigated:
– structure of colorant
– extraction method
– substrate
– light
– painting medium

The influence of the structure of the colorant is illustrated regarding anthraquinones.


The stability of anthraquinones to light decreases with increasing substitution. Brazil-
wood (brazilin) is even less lightfast than the most labile anthraquinone. Madder lake
(vegetable anthraquinone) is more stable than carmine (animal anthraquinone). This
results in a stability order:

Modern alizarin carmine (pure synthetic alizarin) is more stable than


madder lake in laboratory production (natural alizarin, accompanying substances, controlled pro-
duction, and substrate) is more stable than
madder lake (natural alizarin, accompanying substances, variable production, and substrate) is
more stable than
Kermes carmine (natural carminic acid, accompanying substances, variable production, and sub-
strate) is more stable than
Cochineal carmine (natural carminic acid, accompanying substances, variable production, and sub-
strate) is more stable than
Lac dye, brazilwood lake (natural colorant, accompanying substances, variable production, and
substrate).

With flavonoids, it is currently difficult to make an exact statement; however, lakes from
rhamnus berries or quercitron are more stable than woad extract.
Also, the extraction method influences the stability. The decomposition to light is
faster if the dye is extracted from wool (old textiles) than if the colorant is extracted
directly from the original material. A fractional extraction could take place with wool
702 � 7 Paint systems in art

as the competing substrate. There are weaker bonds between colorant and substrate if
wool fibers and mordant are also present in the system. However, the color change in
the painting is the same in both extraction processes.
The nature of the substrate is reflected in the following order of declining stability:

Al(OH)3 is more stable than


Al(OH)3 + CaSO4 or CaCO3 is more stable than
SnO2 .

Lakes on pure aluminum hydroxide substrate are the most stable. Ca complexes of col-
orants are less stable than Al complexes because Ca salts can already compete with the
Al cations during precipitation, forming more stable complexes. After precipitation, pre-
cipitated Ca salts also reduce the amount of more strongly bound Al complexes, and
fillers generally reduce the stability of the lakes.
The influence of light depends on its energy; harmful is mainly high-energy UV radi-
ation below 400 nm wavelength. A UV filter can reduce the light hazard. Unfortunately,
the lead white frequently used in Old Master paintings does not absorb any UV light
and cannot slow down the deterioration of the lakes. In contrast, TiO2 in titanium white
absorbs UV light and provides effective light protection.
The influence of the painting medium is shown in the following order of declining
resistance:

Oil or highly pigmented (water) paint application is