You are on page 1of 20

2

Anterior Eye
JOHN G LAWRENSON

cornea is conventionally divided into four zones (central, para-


Introduction central, peripheral and limbal). The central zone, which covers
A critical aspect of contact lens practice is monitoring the the entrance pupil of the eye, is spherical, approximately 4 mm
ocular response to lens wear, which ranges from acceptable wide, and principally determines high-resolution image for-
physiological changes to adverse pathology. In order to do this, mation on the fovea. The paracentral zone, which lies outside
practitioners must possess a thorough understanding of the the central zone, is flatter and becomes optically important in
normal structure and function of the anterior eye, which is the dim illumination when the pupil dilates. The peripheral zone
subject of this chapter. In the course of reading other chapters is where the cornea is flattest and most aspheric (Klyce et  al.,
in this book, the reader may need to refer back to this chapter 1998). Due to a difference in curvature between its posterior
on the functional anatomy and physiology of the anterior eye and anterior surfaces, the cornea shows a regional variation
in order to develop a fuller understanding of the phenomena in thickness. Centrally the thickness is approximately 0.54
being described.  mm (Doughty and Zaman, 2000), with a peripheral thickness
between 11% and 19% higher than in the centre (Khoramnia
et al., 2007). 
The Cornea
The cornea fulfils two important functions: together with the Microscopic Anatomy
sclera it forms a tough fibrous outer coat that encloses the When the cornea is viewed in transverse section, five distinct
ocular tissues and protects the internal components of the layers can be resolved: epithelium, Bowman’s layer, stroma, Des-
eye from injury. Significantly, the cornea also provides two- cemet’s membrane and endothelium (Fig. 2.1).
thirds of the refractive power of the eye. It is particularly well
suited to its role: the cornea is curved and transparent, and the Epithelium. The epithelium represents approximately 10% of the
air–tear interface provides a refractive surface of good optical thickness of the cornea (55 μm) (Feng and Simpson, 2008). It is a
quality. stratified squamous non-keratinized epithelium, consisting of 5–6
layers of cells (Fig. 2.2), which undergo a constant process of cyclic
CORNEAL ANATOMY
Gross Anatomy
The cornea is elliptical when viewed from in front, with its long
axis in the horizontal meridian (Table 2.1). This asymmetry is
produced by a greater degree of overlap of the peripheral cornea
by opaque limbal tissue in the vertical meridian. The surface
area of the cornea is 1.1 cm2, which represents about 7% of the
surface area of the globe (Maurice, 1984). Topographically, the

TABLE Corneal Dimensions and Related


2.1 Measurements
Parameter Value
Area 1.1 cm2
Diameter
 Horizontal 11.8 mm
 Vertical 10.6 mm
Radius of curvature
  Anterior central 7.8 mm
  Posterior central 6.5 mm
Thickness
 Central 0.54 mm
 Peripheral 0.67 mm
Refractive index 1.376
Power 42 D Fig. 2.1  Transverse section through the cornea. The stroma, which
represents 90% of the thickness of the cornea, is bounded by the epi-
(Data adapted from Bron et al., 1997.) thelium (asterisk) and endothelium (arrow).
10
2  Anterior Eye 11

shedding and replacement to maintain corneal integrity. Three Basal cells consist of single-layer columnar cells with a verti-
distinct epithelial cell types are recognized: a single row of basal cally oriented oval nucleus. Ultrastructurally, they are similar in
cells, 2–3 rows of wing cells and 2–3 layers of superficial (squamous) appearance to wing cells. The plasma membrane similarly shows
cells. In addition, several non-epithelial cells are present (e.g. pronounced infolding and the cytoplasm contains prominent
lymphocytes, macrophages and Langerhans cells). The epithelium intermediate filaments. A variety of cell junctions are present
forms a permeability barrier to water, ions and hydrophilic including: desmosomes, which mediate adhesion between cells;
molecules above a certain size, as well as forming an effective hemidesmosomes, which are involved in the attachment of basal
barrier to the entry of pathogens. Further epithelial specialization cells to the underlying stroma; and gap junctions, which allow for
enhances adhesion between cells, to withstand shearing and intercellular metabolic coupling. Basal cells form the germative
abrasive forces. Furthermore, throughout the thickness of the layer of the cornea, and mitotic cells are often seen at this level. 
epithelium, adjacent cells are connected to one another by water
channels (aquaporins) that are engaged in transcellular water Basal Lamina and Bowman’s Layer. The basal lamina
transport and gap junctions to allow the transfer of ions and small (basement membrane) is synthesized by basal cells. It varies
molecules between cells (Bron et al., 2015). in thickness between 0.5 and 1 μm, and under the electron
Superficial cells are structurally modified for their barrier microscope can be differentiated into an anterior clear zone
function and interaction with the tear film. Scanning elec- (lamina lucida) and a posterior darker zone (lamina densa).
tron microscopy of surface cells shows extensive finger-like The basal lamina is part of a complex adhesion system, which
and ridge-like projections (microvilli and microplicae), which mediates the attachment of the epithelium to the underlying
increase the epithelial surface area. Light, medium and dark stroma (Fig. 2.3). Hemidesmosomes link the cytoskeleton via a
cells can be distinguished depending on the number and pat- series of anchoring fibrils to anchoring plaques in the anterior
tern of surface projections (Pfister, 1973). It has been sug- stroma. The molecular components of this adhesion complex
gested that dark cells, which are relatively free of these surface have been identified and include type VII collagen, integrins,
features, are close to being desquamated into the tear film. By laminin and bullous pemphigoid antigen (Gipson et al., 1987).
contrast, the newly arrived light cells possess a more extensive Bowman’s layer (anterior limiting membrane) varies in thick-
array of surface projections. In high-power transmission elec- ness between 8 and 14 μm. With the light microscope it appears as
tron micrographs, microvilli and microplicae show an extensive an acellular homogeneous zone. Ultrastructurally, it is composed
filamentous covering known as the glycocalyx. The glycocalyx of a randomly oriented array of fine collagen fibrils, which merge
is formed from membrane-bound mucin glycoproteins and is with the fibrils of the anterior stroma (Hogan et al., 1971). Fibrils
important for spreading and attachment of the precorneal tear are composed primarily of collagen types I, III and V. Collagen VII,
film. In accordance with their barrier function, a complex net- associated with anchoring fibrils, is also present. There is evidence
work of tight junctions links superficial cells that exclude water- that Bowman’s layer is formed and maintained primarily by the epi-
soluble dyes such as fluorescein (Bron et al., 2015). thelium, although its function is unclear. The absence of Bowman’s
Wing cells are so named because of their characteristic layer from the cornea of most mammals, and the fact that corneas
shape, with lateral extensions and a concave inferior surface to devoid of this layer over the central cornea following photorefrac-
accommodate the apices of the basal cells. Their nuclei tend to tive keratectomy (PRK) apparently function normally, suggest that
be spherical or elongated in the plane of the cornea. The cell it is not critical to corneal integrity (Wilson and Hong, 2000). 
borders of the polygonal wing cells show prominent infoldings
that interdigitate with adjacent cells, and numerous desmo- Stroma. The stroma is approximately 500 μm thick, and
somes. This arrangement results in a strong intercellular adhe- accounts for 90% of the thickness of the cornea. It is composed
sion. The cytoplasm contains prominent cytoskeletal elements predominantly of collagen fibrils (70% dry weight) embedded in
(predominantly actin and cytokeratin intermediate filaments), a highly hydrated matrix of proteoglycans. A variety of collagen
and although the usual complement of organelles is present they
are few in number.
%DVDOFHOO

'

&6
+'
%/
%RZPDQ¶V
OD\HU


$)

$3 6WURPD
%/
Fig. 2.3  Schematic representation of the adhesion system of the cor-
neal epithelium. Intermediate filaments in the cytoskeleton (CS) are
Fig. 2.2  Corneal epithelium (detail). Three cell types are present: linked through hemidesmosomes (HD) via anchoring fibrils (AF) to an-
basal cells (asterisk), wing cells (arrowhead) and squamous cells (arrow). choring plaques (AP) in the anterior stroma. BL = basal lamina; D = des-
BL = Bowman’s layer. mosome.
12 PART 1  Introduction

Fig. 2.4  Section through the stroma. Keratocytes (arrowed) are locat-
ed between lamellae.
Fig. 2.6  Flat section through the stroma stained with gold chloride.
Keratocytes (arrowed) display a stellate appearance.

physiological measurements of collagen fibre diameter and


spacing can be obtained for the hydrated cornea with the aid of
X-ray diffraction. Using this technique, the mean fibril diameter
in the human cornea is 31 nm, with an interfibrillar spacing of
55 nm (Meek and Leonard, 1993). This narrow fibril diameter
and constant separation, which is a characteristic of corneal
collagen, are necessary prerequisites for transparency.
The interfibrillar space contains a matrix of proteoglycans
(approximately 10% of dry weight). These molecules are highly
sulphated, and along with bound chloride ions create a polyan-
ionic stromal interfibrillar matrix that induces osmotic swelling.
As well as playing a major role in corneal hydration, collagen–
proteoglycan interactions are also thought to be important in
determining the size and spatial arrangement of stromal colla-
gen fibrils (Scott, 1991; Quantock and Young, 2008).
Collagen and proteoglycans are maintained by keratocytes.
These cells occupy 3–5% of stromal volume and lie between col-
lagen lamellae, flattened in the plane of the cornea (Fig. 2.6).
Keratocyte density examined by confocal microscopy and bio-
chemical methods (Møller-Pederson and Ehlers, 1995; Prydal
Fig. 2.5  Electron micrograph of stromal lamellae that cross each other et al., 1998) is non-uniform. Density decreases from superficial
approximately at right angles. Note the regular arrangement of colla- to deep stroma (Hollingsworth et al., 2001) and increases from
gen fibrils within lamellae. centre to periphery. Keratocytes display a large central nucleus
and long slender processes extend from the cell body. Processes
types have been identified. Type I is the major fibril-forming from adjacent cells sometimes make tight junctions with each
collagen, with lesser amounts of types III and V. Non-fibril- other. Cell organelles are not numerous but the usual comple-
forming collagens, including types VI and XII, are found in the ment of organelles, including endoplasmic reticulum, Golgi
interfibrillar matrix (Meek and Boote, 2009). A section taken apparatus and mitochondria, can be observed (Hogan et  al.,
perpendicular to the corneal surface reveals that the collagen 1971).
fibrils are arranged in 200–250 layers (lamellae) running parallel Newer lamellar corneal transplantation techniques have
to the surface (Fig. 2.4). Lamellae are approximately 2 μm thick been developed that allow selective replacement of the diseased
and 9–260 μm wide, and extend from limbus to limbus. Fibrils corneal layers. Deep anterior lamellar keratoplasty (DALK),
of adjacent lamellae make large angles with each other. In the which is increasingly being used to treat keratoconus and cor-
superficial stroma the angles are less than 90°, but fibrils become neal scarring, involves replacement of the affected stroma while
orthogonal in the deeper stroma (Hogan et al., 1971; Meek and retaining the host’s healthy Descemet’s membrane and endo-
Boote, 2009). This preferred orthogonal orientation gradually thelium. Separation between the posterior stroma and Des-
changes in favour of circumferentially aligned collagen at the cemet’s / endothelium can be achieved by intrastromal injection
limbus. This particular arrangement of collagen imparts a high of air, viscoelastic or saline. Dua and co-workers performed a
tensile strength for corneal protection, which is important given histological examination of donor corneas using air bubble sep-
its exposed position. Within lamellae, all collagen fibrils are aration and claimed to have identified a novel ‘pre-Descemet’s
parallel with uniform size and separation (Fig. 2.5). Accurate posterior stromal layer’, which was widely publicized (Dua et al.,
2  Anterior Eye 13

'0

Fig. 2.7  High-powered micrograph of the posterior stroma. Des-


cemet’s membrane (DM) is located between the stroma (S) and the en-
dothelium (arrow).
Fig. 2.9  Tangential (flat) section through the corneal endothelium: a
single layer of polygonal cells with irregular borders can be observed.

N replace damaged or effete cells, there is a progressive reduction


in endothelial cell number with age. At birth the cornea
V contains a total of approximately 500 000 cells, which represents
N
a mean density of 4500 cells / mm2. During infancy, cell loss
G
is particularly marked and a 26% reduction occurs in the first
year of life (Sherrard et  al., 1987). Thereafter the rate of loss
H progressively declines into old age. Since grafted corneas appear
to maintain transparency and functional normality with an
endothelial cell density of less than 1000 cells / mm2, it seems
that normal cell density represents a considerable ‘physiological
Fig. 2.8  Three-dimensional representation of the posterior cornea reserve’ (Klyce and Beuerman, 1998). When viewed en face, for
showing the endothelium (e), Descemet’s membrane (d) and stroma (s). example using a specular microscope, the endothelium appears
A stromal lamella has been reflected to reveal an intralamellar kerato- as a mosaic of polygonal (typically hexagonal) cells (Efron et al.,
cyte (k). 2001). In response to pathology, trauma, age and prolonged
contact lens wear, the endothelial mosaic becomes less regular,
2013). However, the current consensus amongst corneal experts and shows a greater variation in cell size (polymegethism) and
is that this layer is not sufficiently unique to constitute a new shape (pleomorphism) as cells spread to fill gaps caused by
corneal layer (Jester et al., 2013).  cell loss. Under the electron microscope, the lateral borders of
the cells are markedly convoluted and adjacent cells are linked
Descemet’s Membrane. Descemet’s membrane is the basement by tight junctions (with less-frequent gap junctions) (Hogan
membrane of the corneal endothelium. It lies between the et al., 1971). The complement of organelles seen in endothelial
endothelium and the overlying stroma (Fig. 2.7). At birth it is cells reflects their high metabolic activity, with numerous
3–4 μm thick, and increases to a thickness of 10–12 μm in the mitochondria and a prominent rough endoplasmic reticulum. 
adult. In the periphery of aged corneas, Descemet’s membrane
displays periodic sections of thickening, which are known as
Hassall–Henle warts. The anterior one-third of Descemet’s CORNEAL INNERVATION
membrane represents that part produced in fetal life and, under
the electron microscope, is characterized by a periodic banded Source and Distribution of Corneal Nerves
pattern (Fig. 2.8). The posterior two-thirds, which is formed The cornea is the most richly innervated surface tissue in
postnatally, has a more homogeneous granular appearance. the body. Corneal nerves are responsible for the detection of
Descemet’s membrane has a unique biochemical composition somatosensory stimuli and play an important role in initiating
in contrast with other basement membranes (Lawrenson the blink reflex, wound healing and tear secretion (see Sha-
et  al., 1998). The major basement membrane collagen type is heen et al., 2014, for a recent review). The majority of corneal
type IV, whereas in Descemet’s membrane type VIII collagen nerves are sensory and derive from the nasociliary branch of
predominates.  the trigeminal nerve (Ruskell and Lawrenson, 1994). There is
also evidence for the existence of a modest sympathetic inner-
Endothelium. The endothelium is a monolayer of squamous vation from the superior cervical ganglion (Marfurt and Ellis,
cells that lines the posterior surface of the cornea (Fig. 2.9) 1993). Branches from the nasociliary nerve either pass directly
and plays a critical role in maintaining corneal transparency to the eye as long ciliary nerves or traverse the ciliary ganglion,
(Bonanno, 2012). As it has a limited capacity for mitosis to leaving it as short ciliary nerves that enter the eye close to the
14 PART 1  Introduction

concluded that free nerve endings were the exclusive receptors


for pain. Although the specificity theory has subsequently been
challenged, particularly with respect to its exclusivity, the ques-
tion as to whether pain is the only sensory modality perceived
by the cornea remains.
Modern experiments using carefully controlled corneal
stimulation, with a variety of mechanical, chemical and ther-
mal stimuli, have evoked only sensations of irritation or pain.
By contrast, electrophysiological studies of corneal afferent
neurones have identified neurones that respond to mechanical,
thermal and chemical stimulation. However, since the conscious
perception of these sensations has not been demonstrated, it
is likely that such specificity of modality is lost during central
nervous system processing. Electrophysiological recording also
allows for the mapping of receptive fields. These are often large
and overlapping, which explains the inability of the cornea to
localize a stimulus accurately (Belmonte et al., 1997). The sen-
sitivity of the cornea to mechanical stimulation is particularly
acute, and acts as a trigger for the protective blink and lacrimal
reflexes. Cold receptors may be important in signalling evapo-
rative cooling, which is a major determinant of spontaneous eye
blink frequency (Tsubota, 1998).
Fig. 2.10  Whole-mount gold chloride-stained preparation of corneal Corneal afferent fibres also exert important trophic influ-
nerves (arrows) located at basal level. ences. Damage to corneal sensory nerves by surgery, trauma or
infection produces neuroparalytic keratitis – a condition that is
characterized by progressive epithelial cell loss and oedema. The
optic nerve. Nerves destined for the cornea travel initially in mechanism of this trophic role is not fully understood, although
the suprachoroidal space, before crossing the sclera to advance the release of neuropeptides (e.g. substance P and calcitonin
radially towards the cornea. gene-related peptide) may be a factor. Sympathetic nerves also
Most of the 50–80 precorneal nerve trunks, which contain play a role in epithelial maintenance by regulating ion transport
a mixture of myelinated and unmyelinated fibre bundles, enter processes, and cell proliferation and migration during wound
the cornea at mid-stromal level. Myelin is soon lost and the healing. 
unmyelinated nerve fibre bundles divide repeatedly and move
anteriorly to form a rich plexiform network in the anterior one-
third of the stroma. Axons are particularly dense immediately CORNEAL METABOLISM
beneath Bowman’s layer, forming an extensive subepithelial
plexus (Oliveira-Soto and Efron, 2001). From this plexus, axons Source of Oxygen and Nutrients
pass vertically through Bowman’s layer, losing their Schwann In order to perform its vital functions, the cornea requires a
cell sheath in the process. Upon entering the epithelium, axons constant supply of oxygen and other essential metabolites (e.g.
turn through 90° and divide into a series of fine branches that glucose, vitamins and amino acids). However, its avascularity
course between basal cells (Fig. 2.10). Some branches pass into dictates that alternative routes must exist for the provision of its
the more superficial layers before terminating. The density of metabolic needs. There are three possibilities: from the perilim-
nerve terminals is greatest centrally, corresponding to approxi- bal vasculature, from the tear film or from the aqueous humour.
mately 600 terminals / mm2, which results in large overlapping In open-eye conditions the bulk of the oxygen required for the
receptive fields (Shaheen et al., 2014). cornea is obtained from the atmosphere via diffusion across
Corneal nerves display a complex neurochemistry. A variety the precorneal tear film. Under steady-state conditions it can
of neurotransmitters and neuromodulators have been identi- be assumed that the tears are saturated with oxygen, and there-
fied, including acetylcholine, substance P, and calcitonin gene- fore at an oxygen tension corresponding to the atmosphere (155
related peptide. However, it is unclear how these particular mmHg at sea level). It has been estimated that the oxygen ten-
neurochemicals correlate with function (Belmonte et al., 1997).  sion of the aqueous humour in the human eye lies between 30
and 40 mmHg (Klyce and Beuerman, 1998).
Functional Considerations Experiments using nitrogen-filled goggles or sealed scleral
Corneal nerves serve important sensory, reflex and trophic lenses have shown the corneal dependence on tear-side oxygen
functions. Interest in the sensitivity of the cornea dates back to to avoid oedema and maintain normal function. The reason
the 19th century (Lawrenson, 1997), when the pioneering Ger- why the cornea swells during contact lens wear is explained in
man physiologist von Frey concluded that pain was the only Fig. 2.11. During eye closure the oxygen level in the tears is in
sensation perceived by the cornea. This was consistent with his equilibrium with the palpebral vasculature (55 mmHg) (Efron
theory of the specificity of sensory receptors, which maintained and Carney, 1979).
that each sensory modality was subserved by a separate anatom- Significantly, corneal thickness increases by approximately
ically distinct nerve terminal. In his experiments on the cornea, 5% during sleep, and returns to baseline levels within 1 hour
von Frey could elicit only a sensation of pain and, as the cornea of eye opening. It has been suggested that overnight swell-
contained exclusively free (unspecialized) nerve endings, he ing is related to tear film tonicity rather than reduced oxygen
2  Anterior Eye 15

&RUQHDO
6FOHUD HQGRWKHOLXP *O\FRJHQ *OXFRVH
&RUQHDO
&RQWDFW $TXHRXV VWURPD
OHQV KXPRXU &RUQHDO *OXFRVHSKRVSKDWH
+03
&LOLDU\ERG\ HSLWKHOLXP
VKXQW
 1$'3+
&RUQHD ,QFUHDVHG 7HDU
2[\JHQ WKLFNQHVV ILOP
&2 &2
 $73
:DWHU /DFWLF
&U\VWDOOLQH
OHQV DFLG
 2[\J 7&$
HQ /DFWDWH 3\UXYDWH
,ULV F\FOH
 $73
$QWHULRUFKDPEHU
FRQWDLQLQJDTXHRXV
KXPRXU &RQWDFWOHQV
EORFNVR[\JHQ &2
$ %
Fig. 2.11  (A) Cross-section of an eye wearing a contact lens, which Fig. 2.12  Metabolic pathways present in the cornea. HMP = hexose
impedes ingress of oxygen into, and the egress of carbon dioxide from, monophosphate shunt; TCA cycle = tricarboxylic acid (Krebs) cycle;
the cornea. (B) The contact lens blocks oxygen supply to the cornea (1), ATP = adenosine triphosphate; NADPH = nicotinamide adenine dinucle-
causing lactic acid to accumulate in the stroma (2). This draws in water otide phosphate (reduced form).
(3), leading to stromal oedema (4). (Adapted from Efron, N. (1997). Con-
tact lenses and corneal physiology. Biol. Sci. Rev., 9, 29–31.)
diffuse slowly across the endothelium into the anterior cham-
ber. However, during periods of hypoxia the proportion of glu-
availability (Klyce and Beuerman, 1998). The oxygen flux into cose that is metabolized anaerobically increases. The resulting
the cornea can be measured using polarographic oxygen sen- accumulation of lactate causes stromal oedema via an increased
sors. It is in the region of 6 μl / cm2 / h for the cornea as a whole, osmotic load (Klyce, 1981) and localized tissue acidosis (Klyce
although the consumption rate for its composite layers is not and Beuerman, 1998).
equal. Consumption rates have been estimated as 40 : 39 : 21 for The hexose monophosphate shunt (also known as the pen-
the epithelium, stroma and endothelium, respectively (Free- tose phosphate shunt) plays an important role in the corneal
man, 1972). epithelium (Berman, 1981), where it fulfils several important
Several lines of evidence indicate that the aqueous humour is functions, including the generation of intermediates for biosyn-
the primary source of glucose and essential amino acids for the thetic reactions and the prevention of oxidative damage by free
cornea (Maurice, 1984). The glucose concentration of tears is radicals. 
low compared with that in the aqueous humour, and the inser-
tion of nutrient-impermeable implants into the stroma results
CORNEAL TRANSPARENCY
in degeneration of the tissue lying anterior to the implant.
Although exogenous glucose is primarily utilized, glycogen Under normal conditions the cornea is highly transparent,
stores are present in all corneal cells to provide glucose in con- transmitting more than 90% of incident light. Structurally, the
ditions of metabolic stress. cornea is a typical connective tissue consisting principally of a
The role of the perilimbal vasculature in the provision of matrix of collagen and proteoglycans. Under normal circum-
oxygen and nutrients appears limited and it is likely that it is stances such an arrangement would favour light scatter with
significant only for the corneal periphery (Maurice, 1984).  consequent loss of transparency. This raises two fundamental
questions: how is transparency achieved, and how is it main-
Oxidative Metabolism tained? To begin to answer these questions it is necessary to
The cornea derives its energy principally from the oxidative understand the spatial organization of the stromal matrix and
breakdown of carbohydrates (Riley, 1969). Glucose, which the importance of corneal hydration control.
is the primary substrate for the generation of adenosine tri-
phosphate (ATP), is catabolized by three metabolic pathways: Stromal Organization
glycolysis, the tricarboxylic acid (Krebs) cycle and the hex- Maurice (1957) explained the transparency of the cornea on the
ose monophosphate shunt (Fig. 2.12). Anaerobic glycolysis basis of the small diameter and regular separation of the stro-
accounts for the majority (85%) of glucose metabolism. In mal collagen. He suggested that the collagen fibrils of the stroma
this pathway, glucose is first oxidized to pyruvate and then were disposed in a regular crystalline lattice, and that light scat-
subsequently reduced to lactate, with a net yield of two mol- tered by the fibrils is eliminated by destructive interference in
ecules of ATP per mole of glucose. The TCA cycle results in all directions other than the forward direction. This situation
a greater energy yield (36 ATP). This pathway is most active will hold as long as the axes of the collagen fibrils are arranged
in the corneal endothelium, which has the greatest energy in a regular lattice with a separation less than the wavelength of
requirement. light. It has been suggested, however, that the fibrillar arrange-
Metabolic waste products can be potentially damaging if ment need not be in a perfect crystal lattice to maintain trans-
allowed to accumulate. Although carbon dioxide can readily parency (Maurice, 1984), although disruption of short-range
diffuse out of the cornea across its limiting layers, lactate is less order between fibrils will lead to increased scatter and a loss of
easily eliminated. Under normoxic conditions, lactate is able to transparency.
16 PART 1  Introduction

The factors involved in the maintenance of collagen fibril size (QGRWKHOLXP


and spatial order are not fully understood. It has been proposed
that collagen fibril diameters may be controlled by the incorpo-
ration of minor collagens (e.g. type V) into the predominantly
type I fibrils (Meek and Leonard, 1993) and that their spatial
separation is a function of proteoglycan–collagen interactions

'HVFHPHW¶VPHPEUDQH
(Scott, 1991). Proteoglycans are a family of glycoproteins that
consist of a protein core to which are attached sugar chains

$TXHRXV
6WURPD
of repeating disaccharide units termed glycosaminoglycans +2
. 7-
(GAGs). These molecules are increasingly being recognized 1D
as important prerequisites for transparency (Quantock and
Young, 2008; Hassell and Birk, 2010). Proteoglycans were origi- 1D
nally classified according to their glycosaminoglycan composi- + +&2±
tion; however, current nomenclature groups them into families
1D
based on homologous sequences of amino acids in their protein +&2±
core. Corneal stromal proteoglycans are members of the small
&2+2 +&2±+
leucine-rich family, which are small enough to fit in the space
&$
between collagen fibrils. The most prevalent glycosaminoglycan
in the cornea is keratan sulphate, which is found in three types Fig. 2.13  Schematic representation of the corneal endothelial pump.
of proteoglycan: lumican, keratocan and mimecan (Funder- CA = carbonic anhydrase; TJ = tight junction.
burgh et al., 1991; Funderburgh 2000). The other type of corneal
proteoglycan is decorin, which contains a hybrid chondroitin dioxide by the enzyme carbonic anhydrase. Bicarbonate leaves
sulphate / dermatan sulphate side chain. Evidence from several the cell via an apical bicarbonate ion channel. The driving force
sources has shown that lumican and decoran play important for the bicarbonate flux is generated by a sodium–potassium
roles in regulating collagen fibril diameter and maintaining the ATPase that resides on the basolateral endothelial membrane.
spacing between fibrils once formed, which are essential for The energy associated with subsequent sodium re-entry (via
transparency.  Na+ / H+ and Na+ / HCO3− transporters) is coupled to active
HCO3− flux (Hodson et al., 1991).
Hydration Control The epithelium also contributes to corneal hydration con-
The state of corneal hydration is another important determi- trol (Klyce and Beuerman, 1998). The tight junctions between
nant of corneal transparency (Bonanno 2012). The hydrophilic superficial epithelial cells form an effective permeability barrier
properties of the stroma are to a large part determined by pro- to ions and polar solutes. For example, the anionic molecule
teoglycans, which contribute to the fixed negative charge of the sodium fluorescein does not penetrate an intact epithelium.
stroma and produce a passive gel swelling pressure through However, damage to the superficial cells allows fluorescein to
electrostatic repulsion (Hodson, 1997). Physiologically, corneal enter the epithelium, with resulting corneal staining. In addition
hydration is maintained at approximately 78%. If the cornea is to its barrier properties, the epithelium also possesses active ion
allowed to swell ±5% of this value, it begins to scatter significant transport systems for Na+ and Cl−. As these pumps contribute
quantities of light (Hodson, 1997). to the tonicity of the tear film, it is likely that they are involved
Maintenance of physiological corneal hydration is to a large in the maintenance of stromal hydration. 
part dependent on the corneal endothelium, which possesses
both a barrier property and a metabolically driven pump. The Response to Oedema
endothelial barrier to the free passage of molecules from the When corneas swell, light scattering increases, with an ensued
aqueous humour is formed principally by focal tight junctions transparency loss due to the disruption of the regular collagen
between adjacent endothelial cells. However, in contrast to matrix. The collagen fibrils themselves swell very little and most
other barrier epithelia, these junctions are of low electrical resis- of the additional water goes into the interfibrillar spaces. Trans-
tance and allow the passage of ions and small molecules. This mission electron micrographs of oedematous corneas show
leak is offset by the metabolically driven pumping of ions out of fibril aggregation, with the result that large areas are devoid of
the stroma by the endothelium, which maintains a transcellular collagen fibrils (Stiemke et al., 1995). There is evidence that col-
potential difference (aqueous side negative) to balance stromal lagen aggregation occurs as a result of loss of GAGs, which pre-
swelling pressure (Maurice, 1984). Disruption of this osmotic viously had maintained fibre separation (Stiemke et al., 1995). 
gradient will result in stromal fluid imbibition.
The specific endothelial ion transport mechanisms respon-
CORNEAL EPITHELIAL WOUND HEALING
sible for the maintenance of physiological hydration are not
fully understood. A simplified model representing our cur- A smooth and intact corneal epithelium is necessary in order
rent level of knowledge is represented in Fig. 2.13. There is for the cornea to maintain clear vision. However, due to its
compelling evidence that a flux of bicarbonate ions is the pre- exposed position the cornea is potentially vulnerable to a vari-
dominant component of the endothelial ion transport system ety of external insults. It possesses several protective mecha-
(Hodson and Miller, 1976). Subsequent studies have identified nisms to avoid injury, but if tissue damage occurs it is capable
that Cl− transporters may also be important in maintaining the of an effective wound-healing response (Gipson and Inatomi,
pump (Bonanno 2012). The bicarbonate is generated either 1995; Nishida and Tanaka, 1996). Corneal epithelial repair is a
by a Na+ / HCO3− co-transporter located on the basolateral complex process involving an orchestrated interaction between
plasma membrane or via the intercellular conversion of carbon cells and extracellular matrix, which is coordinated by a variety
2  Anterior Eye 17

of growth factors. The process can be divided into three phases: 2UELWDOSDUW
(1) initial covering of the denuded area by cell migration, (2)
cell proliferation to replace lost cells and (3) epithelial differen- 7DUVDO
tiation to re-form the normal stratified epithelial architecture. SDUW
6XSHULRU
Following a full-thickness epithelial defect, fibronectin, an SDOSHEUDO
adhesive glycoprotein, is synthesized and covers the surface of VXOFXV
the bared stroma where it serves as a temporary matrix for cell ,QIHULRU
migration. The adhesion between fibronectin and the epithe- SDOSHEUDO
lium is mediated by integrin–matrix interactions (integrins are a 0DODU VXOFXV
family of cell surface receptors that bind to certain extracellular VXOFXV 1DVRMXJDO
matrix proteins). Several growth factors have been implicated in VXOFXV
the control of the wound-healing response, including epidermal
growth factor, transforming growth factor beta, platelet-derived Fig. 2.14  Surface anatomy of the eyelids. (Adapted from Bron, A. J.,
growth factor and fibroblast growth factor (Gipson and Ina- Tripathi, R. C. & Tripathi, B. (1997). Wolff’s Anatomy of the Eye and Orbit
tomi, 1995). Growth factors, which are produced by a variety of (8th ed.). London: Chapman and Hall.)
sources (e.g. ocular surface epithelia and the lacrimal gland), are
able to regulate the process of epithelial migration, proliferation
and differentiation. There is evidence that epithelial–stromal of the fissure is approximately 30–31 mm, with a vertical height
interactions play an important role in corneal wound healing of 10–11 mm. In the primary position, the upper lid, which is
(Wilson, 2000). Epithelial injury triggers keratocyte apoptosis the larger and more mobile of the two, typically covers approxi-
(programmed cell death) in the anterior stroma, via the release mately the upper third of the cornea, whilst the lower lid is level
of apoptosis-inducing cytokines from epithelial cells. Kerato- with the inferior corneal limbus (Fig. 2.14). Important differ-
cyte apoptosis subsequently triggers a wound-healing cascade, ences in eyelid anatomy exist between Asian and Caucasian eyes
which influences epithelial repair. (Saonanon, 2014). The most obvious feature of the Asian eye is
Regeneration of the corneal epithelium is highly dependent the absent or very low lid fold and fuller upper eyelid.
on the integrity of the limbus (Lavker et al., 2004). Cumulative The eyelid margins are about 2 mm thick from front to back.
evidence indicates that a proportion of limbal basal epithelial The posterior quarter consists of conjunctival mucosa and the
cells possess the properties of stem cells, which are ultimately anterior three-quarters is skin. The junction between the two is
responsible for corneal epithelial replacement. Stem cells have referred to as the mucocutaneous junction. There has recently
several unique characteristics: they are poorly differentiated, been a renewed interest in the marginal zone of the human eyelid,
long lived and have a high capacity for self-renewal. When these with the identification of the role of the inner lid border, termed
cells divide, one of the daughter cells replenishes the stem cell the ‘lid-wiper’ owing to the analogy to a windscreen wiper, in
pool, whilst the other is destined to undergo further cell divi- the distribution of the tear film (Knop et al., 2011a, 2012). Two
sions before differentiating. Such a cell is referred to as a tran- or three rows of eyelashes (cilia) arise from the anterior border
sient amplifying cell. Transient amplifying cells undergo several of the lid margins. These are longer and more numerous in the
rounds of cell division before fully differentiating. These cells upper lid. The lashes receive a rich sensory nerve supply, and
play an important role in epithelial wound healing, where their their sensitivity provides an effective alerting mechanism.
proliferative capacity is increased by shortening cycle times and The meibomian (tarsal) gland orifices emerge just anterior
increasing the number of times that the transient amplifying to the mucocutaneous junction (Fig. 2.15). About 30–40 glands
cells can divide before maturation.  open onto the upper margin, and slightly fewer (20–40) onto
the lower. On eversion of the lids the yellowish meibomian
The Ocular Adnexa acini are visible as yellow clusters through the tarsal conjunc-
tiva (Bron et al., 1991; Knop et al., 2011b). Meibomian glands
The ocular adnexa are those structures that support and pro- can be more clearly visualized using infrared meibography, and
tect the eye, and include the eyelids, conjunctiva and lacrimal instruments that use this method are now commercially avail-
system. They play an important role in the formation of the pre- able (Srinivasan et  al., 2012) (Fig. 2.16). At the medial angle,
ocular tear film and collectively defend the eye against antigenic the eyelid margins enclose a triangular space – the lacus lacri-
challenge. malis – which contains the plica semilunaris and the caruncle.
Lacrimal papillae are small elevations located 5–6 mm from the
EYELIDS medial canthal angle, which contains a small aperture (punc-
tum) that is the opening to the lacrimal drainage system. 
The eyelids are two mobile folds of skin that perform several
important functions: they act as occluders that shield the eyes Muscles of the Eyelids
from excessive light, and through their reflex closure they afford Movements of the eyelids are governed by the coordinated
protection against injury. The lids also form a precorneal tear action of several muscles.
film of uniform thickness during the upturn phase of each
blink. The action of blinking is also important for tear drainage. Orbicularis Oculi. The orbicularis oculi is the sphincter muscle
of the eyelids, and can anatomically be divided into two main
Gross Anatomy divisions: the palpebral and the orbital (Fig. 2.17). Fibres of the
The eyelids are joined at their extremities, termed ‘the canthi’, palpebral division arise from the medial palpebral ligament
and when the eye is open, an elliptical space, the palpebral fis- and arc across the eyelids in a series of half-ellipses, meeting at
sure, is formed between the lid margins. In the adult, the length the lateral canthus to form a lateral raphe. The lateral palpebral
18 PART 1  Introduction

0HLERPLDQRULILFHV G
0XFRXV
PFM
PHPEUDQH
F
0DUJLQDO E
VNLQ
D

/DVKHV D
E
$
F

Fig. 2.17  Schematic representation of the divisions of the orbicularis


oculi and the frontalis. a = pretarsal; b = preseptal; c = orbital; d = fronta-
lis. (Adapted from Bron, A. J., Tripathi, R. C. & Tripathi, B. (1997). Wolff’s
Anatomy of the Eye and Orbit (8th ed.). London: Chapman and Hall.)

ligament also acts as an anchor point. The palpebral division


can be further subdivided into marginal, pretarsal and preseptal
% parts. The marginal part (pars ciliaris), which is also known as
Riolans muscle, is responsible for maintaining the apposition
Fig. 2.15  (A) Schematic representation of the eyelid margin. mcj = mu- of the lid to the cornea during lid closure. A third part of the
cocutaneous junction. (B) Gross appearance of the eyelid margin. muscle (pars lacrimalis) is closely associated with the lacrimal
Openings of the meibomian glands are clearly visible (arrow). (Adapted
from Bron, A. J., Tripathi, R. C. & Tripathi, B. (1997). Wolff’s Anatomy of
outflow pathway. The pars lacrimalis (also known as Horner’s
the Eye and Orbit (8th ed.). London: Chapman and Hall.) muscle) encloses the canaliculi and provides attachments to the
lacrimal sac and its associated fascia.
The orbital part of the orbicularis oculi lies outside the palpe-
bral division and extends for some distance beyond the orbital
margins. Muscle fibres arise predominantly from bone at the
medial orbital rim and appear to sweep around the lids without
interruption as a series of complete ellipses. However, studies
have shown that the muscle fibres of the orbital and palpebral
division of the orbicularis are relatively short (0.4–2.1 mm) and
overlapping (Lander et al., 1996). The regional divisions of the
orbicularis also show a functional distinction. The action of the
palpebral part of the muscle is to produce the reflex or voluntary
closure of the lids during blinking. Contraction of the orbital
division produces the forcible closure of the lids that occurs in
sneezing or in response to a painful stimulus. 

Levator Palpebrae Superioris. The levator palpebrae superioris


is primarily responsible for elevating the upper lid during
blinking and for maintaining an open palpebral aperture. The
levator palpebrae arises from the lesser wing of the sphenoid,
above and anterior to the optic canal, and runs forward along
the roof of the orbit above the superior rectus before terminating
anteriorly in a fan-shaped tendon (aponeurosis). Some fibres
are attached to the anterior surface of the tarsal plate, whilst the
remainder pass between fascicles of the orbicularis (Fig. 2.18).
The superior palpebral sulcus forms at the upper border of the
attachment to the orbicularis. 

Superior and Inferior Tarsal Muscles (of Müller). These


smooth muscles arise from the lower border of the levator in the
upper lid and the inferior rectus in the lower lid, and insert into
the orbital margins of the tarsal plates. The role of the superior
Fig. 2.16  Normal meibomian glands of the upper tarsus (top) and low- tarsal muscle is to assist the levator in maintaining the width
er tarsus (bottom) of a 38-year-old woman, imaged using non-invasive of the palpebral aperture. A mild degree of ptosis results from
infrared meibography. (Image courtesy of Reiko Arita.) damage to its sympathetic nerve supply (Horner’s syndrome). 
2  Anterior Eye 19

D
V
WP
73
(6
W
22F

Fig. 2.18  Diagram showing the relations of the levator palpebrae


superioris. a = levator aponeurosis; tm = superior tarsal muscle (of Mül-
ler); t = tarsal plate; s = orbital septum. (Adapted from Gray, H., Bannis-
ter, L. H., Berry, M. M. & Williams, P. L. (1995) Gray’s Anatomy: The Ana-
tomical Basis of Medicine and Surgery (38th ed.). Edinburgh: Churchill
Livingstone.)

Control of Eyelid Movements 5 ()


Movements of the eyelids occur through the coordinated action
3&
of several muscles – the levator palpebrae, tarsal muscles, the
orbicularis oculi and the frontalis. The elevation of the upper lid
and the control of its vertical position are mediated principally
by the levator. In vertical gaze, lid position and eye movements
are closely linked. During elevation the state of contraction of
Fig. 2.19  Sagittal section through the upper lid. TP = tarsal plate;
the levator is varied to maximize visibility. In extreme upgaze, OOc = orbicularis oculi; R = Riolan’s muscle; EF = eyelash follicles;
lid retraction is augmented by the action of the frontalis, which PC = palpebral conjunctiva; ES = eyelid surface.
elevates the eyebrows. In downgaze, coordinated lid move-
ments similarly occur through levator relaxation. In periodic
and reflex blinks, the levator is spontaneously inhibited prior to They are anchored to the orbital margins by the medial and lat-
orbicularis contraction in lid closure. Similarly, in lid opening eral palpebral ligaments. Each tarsus is approximately 25 mm
the orbicularis relaxes, followed by contraction of the levator. long and 1–2 mm thick. The upper tarsus is semioval with a
Spontaneous eye-blink activity is influenced by both central and height of 11 mm at its midpoint, whereas the inferior tarsus is
peripheral factors (Tsubota, 1998). narrower (4 mm in height). The posterior surface of the eyelid
Compared with the upper lid, the lower lid is relatively is lined by the palpebral conjunctiva, which is firmly adherent
immobile and has no counterpart to the levator palpebrae. The to the underlying tarsal plate. 
depression of the lower lid that occurs in downgaze is due to the
attachment of the sheaths of the inferior oblique and inferior Glands of the Eyelids
rectus muscles to the tarsal plate via a fibrous extension.  Meibomian Glands. The tarsal plates contain the acini and ducts
of the meibomian (tarsal) glands. Ducts are vertically oriented with
Microscopic Anatomy respect to the lid margins, with multiple secretory acini that open
The histological appearance of the upper and lower lids is similar, laterally onto each duct. The glands occupy nearly the full length
and in sagittal section the following six tissue layers can be resolved: and width of each tarsus, and are fewer and shorter in the lower
skin, subcutaneous connective tissue, muscle layer, submuscular lid. Histologically, the acini are lined by a layer of undifferentiated
connective tissue, tarsal plate and palpebral conjunctiva (Fig. 2.19). basal cells that divide, and cells are displaced from the basement
The eyelid skin is thin and very elastic. It is continuous with membrane. As they progress towards the duct they gradually
the palpebral conjunctiva at the lid margin, and keratinization enlarge and develop lipid droplets in their cytoplasm (Fig. 2.20).
is maintained up to this mucocutaneous junction. The subcu- Ultimately, cell membranes rupture and cellular debris, together
taneous connective tissue is composed of a loose areolar tissue with the lipid product, is discharged into the duct.
and contains hair follicles and associated glands. The muscle The stimulus for meibomian gland secretion is unclear.
layer consists of striated muscle fibres of the orbicularis oculi, Although a modest autonomic innervation of the meibomian
which are arranged in bundles (fascicles) separated by con- glands has been demonstrated, there is still some doubt regard-
nective tissue. The orbicularis extends throughout the lid. The ing a neuromodulation of glandular secretion; it is likely that the
marginal part of the muscle (Riolan’s muscle) is separated from principal control of the glands is hormonal, and both androgens
the pretarsal portion by connective tissue that contains the eye- and oestrogens have been shown to regulate meibomian secre-
lash follicles. The loose submuscular connective tissue layer lies tion (Sullivan et  al., 2000; Knop et  al., 2011b). A long ductal
between the orbicularis and the tarsal plate and contains the system carries the secretion to the lid margin, and the compres-
major nerves and vessels of the lid. sive action of the palpebral division of the orbicularis oculi on
The tarsal plates (tarsi) are composed of dense fibrous con- the meibomian ducts facilitates the flow of lipid and its eventual
nective tissue and provide support and determine lid shape. delivery onto the lid margins. 
20 PART 1  Introduction

2
%


/
'

Fig. 2.22  Schematic representation of a mid-sagittal section through


the eyelid and conjunctival sac showing the different conjunctival re-
Fig. 2.20  Histological section showing meibomian gland acini. Secre- gions. M = marginal; T = tarsal; O = orbital; B = bulbar; L = limbal; F = for-
tory cells degenerate (asterisk) as they approach the duct (D). nical.



=

0

() 0 

Fig. 2.23  Static dimensions of the conjunctival sac in millimetres.


M = medial canthus. (Adapted from Ehlers, N. (1965). On the size of the
conjunctival sac. Acta Ophthalmol., 43, 205–210.)
Fig. 2.21  Histological section through the ciliary zone of the eyelid.
Glands of Zeis (Z) discharge their contents into an eyelash follicle (EF), palpebral conjunctiva. Veins of the eyelids empty into veins of
which contains the remnants of an eyelash. M = gland of Moll. the forehead and temple, and some empty into the ophthalmic
vein. Lymphatics drain to the preauricular and submandibular
Glands of Zeis and Moll. Ciliary glands of Zeis and Moll are lymph nodes. 
found in association with eyelash follicles (Takahashi et  al.,
2013) (Fig. 2.21). Zeis glands are unilobular sebaceous glands
that open directly into the follicle. The function of their oily THE CONJUNCTIVA
secretion is to lubricate the lashes to prevent them from drying
out and becoming brittle. Glands of Moll are modified sweat Gross Anatomy
glands (apocrine) consisting of an unbranched spiral tubule. The conjunctiva is a thin transparent mucous membrane that
The exact function of these glands is unclear, although their extends from the eyelid margins anteriorly, providing a lining to
secretion is rich in IgA, which suggest a role in the immune the lids, before turning sharply upon itself to form the fornices,
defence of the ocular surface (Stoeckelhuber et al., 2003).  from where it is reflected onto the globe, covering the sclera up
to its junction with the cornea. It thus forms a sac that opens
Blood and Nerve Supply anteriorly through the palpebral fissure. The conjunctiva is con-
Nerves of the Eyelids. The levator palpebrae and orbicularis ventionally divided into the following regions: marginal, tarsal,
oculi muscles are innervated by the oculomotor and facial orbital (these three collectively form the palpebral conjunctiva),
nerves, respectively. The sensory supply of the upper lid bulbar and limbal (Fig. 2.22).
derives from branches of the ophthalmic nerve (supraorbital, The static dimensions of the conjunctival sac in the primary
supratrochlear and lacrimal). The supply to the lower lid comes position are illustrated in Fig. 2.23 (Ehlers, 1965). The marginal
from branches of the maxillary nerve (zygomatic, infraorbital).  zone extends from a line immediately posterior to the openings
of the tarsal glands and passes around the eyelid margin, from
Blood and Lymphatic Supply to the Eyelids. The arterial where it continues on the inner surface of the lid as far as the
supply derives from branches of the ophthalmic, lacrimal subtarsal fold (a shallow groove that marks the marginal edge
and infraorbital arteries, which contribute to two palpebral of the tarsal plate). The tarsal conjunctiva is highly vascular
arcades in the upper lid and one in the lower. Branches from and is firmly attached to the underlying fibrous connective tis-
these arcades supply the skin, orbicularis, tarsal glands and sue. From the convex border of the tarsal plate, the orbital zone
2  Anterior Eye 21

Fig. 2.24  High-power slit-lamp view of the conjunctival palisades of


Vogt (asterisks) at the lower limbus.

extends as far as the fornices. Over this region the conjunctiva is


more loosely attached to underlying tissues, and so readily folds.
Fig. 2.25  Histological section through the bulbar conjunctiva. E = epi-
Elevations of the conjunctival surface in the form of papillae thelium; S = stroma. Goblet cells can be seen in the epithelium (arrows).
and lymphoid follicles are commonly observed in this region. The stroma can be resolved into an adenoid layer (arrowhead) and a
The transparency of the bulbar conjunctiva readily permits deep fibrous layer (asterisk).
the visualization of conjunctival and episcleral blood vessels.
Here, the conjunctiva is freely movable owing to its loose attach-
ment to Tenon’s capsule (the fascial sheath of the globe). As the
bulbar conjunctiva approaches the cornea, its surface becomes
smoother and its attachment to the sclera increases. The limbal
conjunctiva extends approximately 1–1.5 mm around the cornea.
Its junction with the cornea is ill defined, particularly in the ver-
tical meridian, owing to a variable degree of conjunctival / scleral
overlap. The limbus has a rich blood supply, and in the majority
of individuals a radial array of connective tissue elevations – the
palisades of Vogt – can be seen adjacent to the corneal margin
(Fig. 2.24). The palisades are most prominent in the vertical
meridian, and their visibility is enhanced in pigmented eyes.  Fig. 2.26  Histological section through the palisade region. Connec-
tive tissue ridges can be seen projecting into the overlying epithelium
Microscopic Anatomy (arrows).
In histological section, two distinct layers can be resolved: an
epithelium containing a variable number of goblet cells, and a unique array of connective tissue ridges (the palisades of Vogt),
vascular stroma that consists of a superficial lymphoid layer and which project into the overlying epithelium (Fig. 2.26). Clinical
a deep fibrous layer (Fig. 2.25). The appearance of the conjunc- and experimental evidence suggests that the palisades are the
tiva shows a marked regional variability. repositories of stem cells and therefore act as the regenerative
organ of the corneal epithelium (Dua and Azuara-Blanco,
Epithelium. In the marginal zone, the epithelium is stratified 2000). The conjunctival epithelium additionally contains several
and squamous with relatively few goblet cells, although this has non-native cell types, including dendritic cells, melanocytes and
recently been disputed following the description of intracellular lymphocytes. 
crypts lined with goblet cells lying deep within the epithelium in
the region of the so-called ‘lid wiper’ region (Knop et al., 2012). Goblet and Other Secretory Cells. Goblet cells provide the
It has been suggested that a subpopulation of epithelial cells that mucous component of the tear film. They arise in the basal
lie close to the mucocutaneous junction may be acting as stem cell layers and migrate to the surface, there becoming fully
cells for the palpebral conjunctiva (Wirtschafter et  al., 1999). differentiated. Mature goblet cells are larger than the surrounding
Approaching the tarsus, the epithelium thins to 2–3 layers of epithelial cells and contain a peripherally placed nucleus. The
cuboidal cells with scattered goblet cells. The epithelium of the cytoplasm is packed with membrane-bound secretory granules
orbital zone is slightly thicker (2–4 cells) with more numerous that discharge from the apical surface in an apocrine manner.
goblet cells. The number of goblet cells declines over the bulbar The number of goblet cells shows a marked regional variation
conjunctiva and at the limbus the epithelium is again stratified in density (Kessing, 1968) (Fig. 2.27), and these cells are
squamous, and goblet cells are absent. The limbus contains a occasionally seen lining intraepithelial crypts (of Henle).
22 PART 1  Introduction

Fig. 2.27  Diagram showing the regional variation in goblet cell densi-
ty. Goblet cell density is greatest over the caruncle, plica semilunaris and
inferior nasal palpebral conjunctiva. (Adapted from Bron, A. J., Tripathi,
R. C. & Tripathi, B. (1997). Wolff’s Anatomy of the Eye and Orbit (8th ed.).
London: Chapman and Hall. Reproduced from Bron, 1997.)
Fig. 2.28  Histological section through a lymphoid follicle (F). Note the
modification of the overlying epithelium (asterisk).
The apices of many surface epithelial cells of the conjunc-
tiva contain numerous carbohydrate-containing secretory
vesicles, which are seen to migrate to the cell surface where
they fuse with the plasma membrane (Dilly, 1985). It is likely
that this represents a mechanism for recycling the cell sur-
face glycocalyx rather than a secondary source of secretory
mucin. 

Conjunctival Stroma. The conjunctival stroma (substantia


propria) is variable in thickness. It can be resolved into
two distinct layers: a superficial adenoid layer and a deeper
fibrous layer (see Fig. 2.25). The adenoid layer contains
numerous lymphocytes with local accumulations in the form
of lymphoid follicles (Fig. 2.28). Follicles represent aggregates
of predominantly B cells, which form part of the so-called
conjunctiva-associated lymphoid tissue (Knop and Knop, Fig. 2.29  High-power slit-lamp photograph showing the limbal vascu-
2005). The adenoid layer also contains a large number of mast lar arcades. (Courtesy of Eric Papas.)
cells, which play a major role in ocular allergy (McGill et  al.,
1998). The deep fibrous layer is generally thicker than the
adenoid layer and contains the majority of conjunctival blood Blood Vessels and Lymphatics. The arterial supply derives
vessels and nerves.  from two sources: palpebral branches of the nasal and lacrimal
arteries, and anterior ciliary arteries.
Innervation and Blood Supply Palpebral vessels serve two vascular arcades within the eye-
Nerves. The conjunctiva receives nerves from sensory, lid. The inferior (marginal) arcade sends branches through the
sympathetic and parasympathetic sources. Sensory nerves, tarsal plate to the eyelid margin and tarsal conjunctiva. The
which are trigeminal in origin, reach the conjunctiva via superior (palpebral) arcade supplies the tarsal, orbital, fornix
branches of the ophthalmic nerve. The principal function and bulbar conjunctiva. The limbal zone, in contrast, is served
of these fibres is to equip the conjunctiva with the ability by anterior ciliary arteries. The anterior ciliary arteries travel
to detect a variety of sensations – for example, touch, pain, along the tendons of the rectus muscles and give off branches
warmth and cold. Sensory nerve terminals include both at episcleral level prior to dipping down into the sclera to link
free (unspecialized) nerve endings and the more complex with the major iridic circle. Episcleral branches pass forward
corpuscular endings (classically referred to as Krause end and loop back a few millimetres short of the cornea to become
bulbs) (Lawrenson and Ruskell, 1991). Conjunctival blood conjunctival vessels. Forward extensions of these vessels form
vessels receive a dual autonomic innervation. Parasympathetic the limbal arcades (limbal loops), which are a complex net-
fibres issuing from the pterygopalatine ganglion and work of fine capillaries (Fig. 2.29). Conjunctival veins are more
sympathetic fibres from the superior cervical ganglion numerous than arteries. They can be readily differentiated from
are responsible for vasodilation and vasoconstriction, arteries owing to their larger calibre, darker colour and more
respectively.  tortuous path. 
2  Anterior Eye 23

2'

3'

/$


Fig. 2.31  Low-power light micrograph of the lacrimal gland. Acini are
arrowed. Adipose connective tissue (asterisks) extends across the gland.

Fig. 2.30  Lateral view of the orbit showing the position of the lacrimal
gland. The levator aponeurosis (LA) partially divides the gland into an
orbital (OD) and palpebral (PD) division. (Adapted from Kronfeld, P. C.,
McHugh, S. L. & Polyak, S. L. (1943). The Human Eye in Anatomical
Transparencies. Rochester, NY: Bausch & Lomb.)

Functional Considerations
The conjunctiva contributes the mucin component of the pre-
ocular tear film and plays an important role in the defence of
the ocular surface against microbial infection. Mucins are a
family of high-molecular-weight glycoproteins that include
membrane-bound and secretory varieties (Corfield et al., 1997;
Gipson and Inatomi, 1997; Hodges and Dartt, 2013). Goblet
cells are the primary source of secretory mucin, whilst sur-
face epithelial cells of both the conjunctiva and cornea possess
mucin-like molecules within their glycocalyx. The conjunctiva Fig. 2.32  Electron micrograph of part of a lacrimal acinus showing
also forms part of a common mucosal defence system, which light and dark secretory cells.
is an important component of the defence of the human body
against microorganisms (McClellan, 1997; Knop and Knop, a lower palpebral lobe, which can often be visualized through
2005). The conjunctiva possesses the immunological capacity the conjunctiva upon lid eversion (Bron, 1986). The gland is
for antigen processing, and cell-mediated and humoral immu- pinkish in colour, with a lobulated surface. Between 6 and 12
nity. Humoral immunity is provided by specific antibodies (par- ducts leave the gland through the palpebral lobe and discharge
ticularly immunoglobulin A [IgA]) produced by transformed B into the conjunctival sac at the upper lateral fornix. 
cells (plasma cells) in the stroma. T lymphocytes form the basis
of cell-mediated immunity.  Microscopic Anatomy. The lacrimal gland is tubuloacinar in
form (Fig. 2.31). Its secretory units (acini) contain secretory
cells surrounded by myoepithelial cells (Ruskell, 1975). Acinar
LACRIMAL SYSTEM
secretory cells show extensive folding of their plasma membrane
The lacrimal apparatus provides for the production and main- and apical microvilli. Adjacent cells are linked by tight junctions
tenance of the preocular tear film. The normal function of this that restrict diffusion between cells. The most prominent feature
system is essential for the integrity of the ocular surface and the of these cells is the presence of abundant secretory granules.
provision of a smooth refractive surface. The lacrimal apparatus Two principal secretory cell subtypes have been identified on
comprises a secretory system that includes the main and acces- the basis of their granule content (Fig. 2.32). The majority of
sory lacrimal glands, and a drainage system that consists of the cells contain dark granules (dark cells), with a smaller number
paired puncta and canaliculi, the lacrimal sac and the nasolac- of cells containing light granules (light cells). The functional
rimal duct. significance of this heterogeneity is uncertain at present. Ducts
consist of a single layer of cuboidal cells that lack secretory
Lacrimal Gland granules. Myoepithelial cells are dendritic cells that are closely
Gross Anatomy. The main lacrimal gland is the key provider associated with the perimeter of acini and ducts. It is likely that
of the aqueous component of the tears. The gland is located in a these contractile cells play a role in the expulsion of tears from the
shallow depression of the frontal bone behind the superolateral gland. The interstices of the gland contain numerous blood vessels
orbital rim (Fig. 2.30). It is partially split by the aponeurosis and nerves. A large population of immune cells (particularly IgA-
of the levator palpebrae into an upper larger orbital lobe and secreting plasma cells) are also found between acini. 
24 PART 1  Introduction

6&

6& /DFULPDOJODQG

$QWLJHQ
,J$ /6

,J$
6
&,
J$
3

*, &
&LUFXODWLRQ
WUDFW

3H\HU¶VSDWFKHV 1/'

Fig. 2.33  Diagram showing the role of the gastrointestinal tract gen-
erating specific immunoglobulin A (IgA) in the lacrimal gland. Antigens
which challenge the ocular surface ultimately drain to the gastrointes-
tinal (GI) tract where they stimulate B cells in Peyer’s patches (gut-as- Fig. 2.34  Illustration of the lacrimal drainage system. C = canaliculi;
sociated lymphoid tissue). Sensitized B cells then pass to the lacrimal LS = lacrimal sac; P = punctum; NLD = nasolacrimal duct. (Adapted from
gland via the circulation. SC = secretory component. (Adapted from Al- Kronfeld, P. C., McHugh, S. L. & Polyak, S. L. (1943). The Human Eye in
lansmith, M. R. (1992). The Eye and Immunology. Maryland Heights, MO: Anatomical Transparencies. Rochester, NY: Bausch & Lomb.)
Mosby. Copyright Elsevier 2002.)

It has been demonstrated that the lacrimal gland also


Blood and Nerve Supply. The arterial supply to the lacrimal secretes into the tears growth factors that are important for the
gland is provided by the lacrimal artery, which enters the posterior maintenance of the ocular surface and epithelial wound healing
border of the gland. Venous drainage occurs via the lacrimal (Pflugfelder, 1998). Prominent amongst these growth factors are
vein. A rich autonomic innervation includes secretomotor epidermal growth factor and transforming growth factor beta. 
(parasympathetic) fibres that issue from the pterygopalatine
ganglion and sympathetic (vasomotor) fibres from the carotid Lacrimal Drainage System
plexus. The lacrimal nerve traverses the gland to provide a sensory Tears collect at the medial canthal angle, where they drain into
innervation to the conjunctiva and lateral aspect of the eyelid.  the puncta of the upper and lower lids. Each punctum is a small
oval opening approximately 0.3 mm in diameter that is located
Accessory Lacrimal Glands. Numerous small accessory at the summit of an elevated papilla. From each punctum the
lacrimal glands, which include the eponymous glands of canaliculus passes first vertically for about 2 mm and then turns
Wolfring and Krause, are found within the conjunctival stroma. sharply to run medially for about 8 mm (Fig. 2.34). At the angle,
They have a particular predilection for the upper fornix and a slight dilation, the ampulla, can be seen. The canaliculi con-
above the tarsal plate, and, on the basis of proportion of total verge towards the lacrimal sac, usually forming a common can-
lacrimal tissue, it has been estimated that they contribute aliculus before entry. The lacrimal sac occupies a fossa formed
5–10% of aqueous tear volume. Structurally, they have a similar by the maxillary and lacrimal bones. It measures 1.5–2.5 mm in
appearance to the lacrimal gland proper. However, true acini are diameter and approximately 12–15 mm in vertical length. From
absent and glands consist of elongated tubules that connect with the lacrimal sac tears drain into the nasolacrimal duct, which
ducts opening onto the conjunctival surface (Seifert et al., 1993).  extends for about 15 mm, passing through a bony canal in the
maxillary bone, to an opening in the nose beneath the inferior
Functional Considerations. In addition to its role as the principal nasal turbinate. A fold of mucosa is often observed at the ter-
provider of the aqueous phase of the tear film, the lacrimal gland mination of the duct: this has been termed ‘the valve of Hasner’,
is also a major component of the ocular sensory immune system, although there is no strong evidence that it functions as a valve.
which acts as the first line of defence against microbial infection The process of tear drainage is an active process mediated
(Sullivan and Sato, 1994). The secretory immune system is by the contraction of the orbicularis during blinking (Doane,
mediated through secretory IgA. The lacrimal gland is the main 1981). Tears enter the canaliculi principally by capillary action.
source of tear IgA and the gland contains a large number of IgA- During the early part of the blink the puncta are occluded as
producing plasma cells. The mechanism by which an antigenic the orbicularis further contracts. The canaliculi and lacrimal
challenge of the ocular surface induces a lacrimal antibody sac are also compressed, forcing fluid into the nose. An alterna-
response is not fully understood. However, as the administration tive hypothesis has been proposed whereby orbicularis contrac-
of an antigen by a gastrointestinal route raises specific IgA levels tion dilates the sac, creating a negative pressure, which draws
in tears, one suggested mechanism is that ocular antigens – after in the tears from the canaliculi (Jones, 1961). An investigation
drainage through the nasolacrimal duct – stimulate B cells in by Paulsen et al. (2000) described a vascular plexus embedded
gut Peyer’s patches. These sensitized B cells then populate the in the wall of the lacrimal sac and duct that may influence tear
lacrimal gland where they transform into plasma cells (Fig. 2.33). outflow. It is postulated that opening and closing of the lumen of
2  Anterior Eye 25

TABLE Physical Properties of the Preocular Tear


2.2 Film
Parameter Value 3URWHLQ
LRQV
Osmolarity 302 (±6.3) mOsm/l + 2
pH 7.45 3URWHLQ
Thickness 3 μl LRQV
+2
Volume 7.0 (±0.2) μl &RQMXQFWLYDO
Rate of production HSLWKHOLXP
 Unstimulated 1–2 μl / min
 Stimulated >100 μl / min /DFULPDO
Refractive index 1.336 JODQG
$FFHVVRU\
ODFULPDO
JODQGV
the lacrimal passages can be achieved by regulating blood flow
within this plexus. 

The Preocular Tear Film

7HDUILOP
FUNCTION AND PROPERTIES OF THE 0HLERPLDQ
PREOCULAR TEAR FILM JODQGV

The tear film is a complex fluid that covers the exposed parts of the
ocular surface framed by the eyelid margins. The physical charac-
teristics of this fluid are summarized in Table 2.2. Classically, the
(\HOLG
tear film has been regarded as a trilaminar structure with a super-
ficial lipid layer secreted by the meibomian glands, which overlies
an aqueous phase derived from the main and accessory lacrimal
glands, and an inner mucinous layer consisting of membrane- ,RQV
spanning mucins of the ocular surface epithelium and secretory +2
/LSLG
mucins produced mainly by conjunctival goblet cells. The tear
film performs several important functions, which can be broadly
classified as optical, metabolic support, protective and lubrication.
By smoothing out irregularities of the corneal epithelium, the &RQMXQFWLYDO
tear film creates an even surface of good optical quality that is re- 0XFXV HSLWKHOLXP
*REOHWFHOOV
formed with each blink. The air–tear interface forms the principal
refractive surface of the optical system of the eye and provides two-
thirds (43 D) of its total refractive power. As the cornea is avascular Fig. 2.35  Schematic representation of the orbital glands, which contrib-
ute the various components of the preocular tear film. (Adapted from Dartt,
it is dependent on the tear film for its oxygen provision. When the D. A. (1992). Physiology of tear production. In M. A. Lemp & R. Marquardt
eye is open the tear film is in a state of equilibrium with the oxygen (eds) The Dry Eye: A Comprehensive Guide. Berlin: Springer-Verlag.)
in the atmosphere, and gaseous exchange takes place across the
tear interface. The constant turnover of the tear film also provides
a mechanism for the removal of metabolic waste products. (i.e. in response to strong physical or emotional stimulation) is
Tears play a major role in the defence of the eye against mediated by the main lacrimal gland. However, Jordan and Baum
microbial colonization. The washing action of the tear fluid (1980) questioned the concept of basic and reflex secretion, and
reduces the likelihood of microbial adhesion to the ocular sur- suggested that it is more accurate to think of tear output as a con-
face. Moreover, the tears contain a host of protective antimicro- tinuum, whereby the rate of production is proportional to the
bial proteins. The tear film acts as a lubricant, smoothing the degree of sensory or emotive stimulation (Dartt, 2009). This con-
passage of the lids over the corneal surface and preventing the cept would also mean that a functional distinction between main
transmission of damaging shearing forces. To facilitate this, tear and accessory lacrimal glands, in terms of basal and reflex tear
fluid displays non-Newtonian behaviour with respect to shear production, is unnecessary. Rather, it is more likely that tear flow
(Tiffany, 1991). Newtonian fluids maintain a constant viscosity is the combination of contributions from both glands, although
with increasing shear rates. By contrast, tear fluid has a rela- the output from the accessory glands alone is sufficient to main-
tively high viscosity between blinks to aid stability, and with tain a stable tear layer (Maitchouk et al., 2000). 
increasing shear rates during the blink process the viscosity falls
dramatically, thereby easing the movement of the lids over the SOURCES AND COMPOSITION
ocular surface.
Tears are composed of a complex secretion that combines the
Tear Production products of several glands (Fig. 2.35). Although the precise com-
Jones (1966) first used the terms ‘basic (basal)’ and ‘reflex’ to position of tear fluid varies with collection method, flow rate and
describe tear flow. He proposed that the accessory lacrimal glands time of day, it can be considered as a watery secretion containing
were the basic (minimal flow) secretors, and that reflex secretion electrolytes and proteins, with lesser amounts of lipid and mucin.
26 PART 1  Introduction

is a constitutively secreted lacrimal protein whose rate of secre-


TABLE Biochemical Composition of the Preocular tion is independent of flow rate. During sleep, the levels of IgA
2.3 Tear Film increase as secretory IgA production continues and as acinar
Component Concentration secretion declines (Sack et  al., 1992). IgA plays an important
role in the defence of the ocular surface against microbial infec-
ELECTROLYTES*
tion by preventing bacterial and viral adhesion, and inactivating
Na+ 135 mEq / l
Cl− 131 mEq / l
bacterial toxins. Other immunoglobulins (e.g. IgG and IgM) are
K+ 36 mEq / l present in tears at much lower levels.
HCO3− 26 mEq / l Lysozyme, lactoferrin and lipocalin, in contrast, originate
Ca2+ 0.46 mEq / l from acinar cells and their rate of secretion roughly matches
Mg2+ 0.36 mEq / l flow rate. Lysozyme is a well-known bacteriolytic protein that
MAJOR PROTEINS* has the ability to lyse the cell wall of several Gram-positive bac-
Lysozyme 2.07 g / l teria. Lactoferrin serves an important bacteriostatic function by
Secretory IgA 3.69 g / l binding iron and making it unavailable for bacterial metabo-
Lactoferrin 1.65 g / l lism. It also acts as a free radical scavenger, thereby reducing
Lipocalin 1.55 g / l
Albumin 0.04 g / l free-radical-mediated cell damage (Tiffany, 1997). Lipocalins
IgG 0.004 g / l are a family of lipid-binding proteins with an affinity for a broad
LIPIDS† array of lipids, including fatty acids, phospholipids and choles-
Wax esters 41% terol. It has been suggested that tear lipocalins act as scavengers
Cholesteryl esters 27.3% for a wide range of meibomian lipids, which could spill onto
Polar lipids 14.8% the corneal surface and perturb its wettability (Glasgow et al.,
Hydrocarbons 7.5% 2000). Furthermore, lipocalin may promote lipid solubility at
Diesters 7.7% the aqueous–lipid interface to facilitate the formation of a thin
Triacylglycerides 3.7%
Fatty acids 2.0% layer of lipid on the surface of the tear film. 
Free sterols 1.6%
Mucins
MUCIN‡
MUC1 nd
Mucins are a family of high-molecular-weight glycoproteins, of
MUC5AC nd which sugars contribute up to 85% of their dry weight. Structur-
MUC4 nd ally, they consist of a polypeptide backbone to which chains of
MUC16 nd sugar molecules attach via O-linkages to the amino acids serine
(Data adapted from Tiffany, 1997.) and threonine. Mucins are a heterogeneous group of molecules
Sources: that can be subdivided into secretory and integrated-membrane
*Main and accessory lacrimal glands. varieties (Corfield et al., 1997; Hodges and Dartt, 2013). So far,
†Meibomian glands.
‡Epithelial cells / goblet cells.
modern molecular biology techniques have identified up to
nd  = not determined.
20 mucin (MUC) genes, although only four of these (MUC1,
MUC5AC, MUC4 and MUC16) are expressed on the human
ocular surface (Gipson and Inatomi, 1997; McKenzie et al., 2000;
Electrolytes Pflugfelder et al., 2000; Mantelli and Argüesco, 2008). The epithelia
Human tears contain approximately the same range of electro- of the cornea and conjunctiva express the transmembrane mucins
lytes as found in plasma (Tiffany, 1997). Table 2.3 gives typi- MUC1, and to a lesser extent MUC4 and MUC16, which attach
cal values for the ionic composition of human tears. However, to apical microvilli where they form a hydrophilic base to facilitate
as the electrolyte content of tears varies with flow rate, there is the spreading of the goblet-cell-derived mucin MUC5AC. Mucins
significant variation in measured values. During the process play a major role in stabilizing and spreading the tear film and
of secretion by the lacrimal gland, there is a process of active provide protection against desiccation and microbial invasion
electrolyte transport that is coupled to the passive movement of (Gipson and Inatomi, 1997; Hodges and Dartt, 2013). 
water by an osmotic process. Acinar-derived fluid is essentially
an isotonic ultrafiltrate of plasma. Its composition is altered as Lipids
it passes along the ductal system, where further chloride and The source of lipids in the tear film is the meibomian glands
potassium ions are secreted. A variety of ion transport proteins embedded within the tarsal plates of each lid. The blinking pro-
have been identified in acinar cells, including sodium–potas- cess is an important mechanism in the expulsion of the secre-
sium ATPase and potassium and chloride channels.  tion from the glands (Tiffany, 1995). Meibomian lipid (also
known as meibum) is delivered directly as a clear oil onto the lid
Proteins margins and is spread over the tear film from the inner edge of
Tear proteins are thought to originate from three main sources: the lid margins with each blink. The thickness of the lipid layer
the lacrimal gland, ocular surface epithelia and conjuncti- is variable (mean thickness 42 nm, range 15–157 nm; King-
val blood vessels. The major lacrimal proteins include secre- Smith et al., 2000, 2010), and depending on thickness gives rise
tory IgA, lysozyme, lactoferrin and lipocalin (formerly known to characteristic interference patterns when viewed in specular
as tear-specific prealbumin) (see Table 2.3). IgA, which is reflection (Fig. 2.36) (Guillon, 1998). Meibomian secretion con-
the major immunoglobulin in tears, is secreted as a dimer by sists of a complex mixture of lipids (Table 2.3), including wax
plasma cells in the interstices between lacrimal acini. It then and cholesteryl esters (which together constitute approximately
binds to a receptor on the basolateral aspect of acinar cells, and 70% of meibum), fatty acids and fatty alcohols (Tiffany, 1995;
is transcytosed across the cell and secreted into tear fluid. IgA Butovich, 2013). The primary functions of this secretion are to
2  Anterior Eye 27

$LU
+\GURSKRELFOLSLG
+\GURSKLOLFOLSLG
/LSLG $TXHRXV

$TXHRXVSKDVH
6HFUHWHGPXFLQ
08&$&

0XFLQRXVSKDVH *O\FRFDO\[
08&

0LFURYLOOL
(SLWKHOLXP
Fig. 2.37  Diagram showing the composition of the preocular tear
Fig. 2.36  Lipid layer of the preocular tear film viewed in specular re- film. Insets show details of the glycocalyx and lipid–aqueous interface.
flection. A ‘wave’ appearance can be seen, which represents the most (Adapted from Corfield, A. P., Carrington, S. D., Hicks, S. J. et al. (1997).
commonly observed lipid pattern in the population. Ocular mucins: purification, metabolism and functions. Prog. Retin. Eye
Res., 16, 627–656.)

provide a hydrophobic barrier at the lid margin to prevent over- is thought to consist of a mixture of soluble and gel-forming
spill of tears, and to cover the surface of the tear film to retard mucins (Hodges and Dartt, 2013). 
evaporation (Craig and Tomlinson, 1997). 
Conclusion
MODELS OF TEAR FILM STRUCTURE
It is clear from the above account that our understanding of the
The classical trilaminar model of tear film structure in terms of structure and function of the anterior eye is far from complete,
a superficial lipid layer, a middle aqueous layer and deep mucin which places certain limits on our understanding of clinical,
layer, first proposed by Wolff and subsequently modified by contact-lens-related phenomena. It is essential, therefore, that
Holly and Lemp (1977), has received broad acceptance. How- future research continues to focus on fundamental aspects of
ever, the results of recent studies have led to a re-evaluation of ocular anatomy and physiology, as well as on the more applied
the nature of the aqueous and mucinous layers. Several pieces of clinical applications that are described in the remainder of this
evidence have suggested that the mucin contribution to the tear book.
film is much greater than was previously thought (Prydal et al.,
1992), and an alternative tear film model, which possesses a Access the complete references list online at
substantial mucinous phase, has been proposed (Fig. 2.37). The http://www.expertconsult.com.
nature of the mucinous phase has not been fully established, but
REFERENCES
Allansmith, M. R. (1992). The Eye and Immunol- Funderburgh, J. L. (2000). Keratan sulphate: struc- King-Smith, P. E., Hinel, E. A., & Nichols, J. J.
ogy. St Louis, MO: Mosby. ture, biosynthesis and function. Glycobiology, 10, (2010). Application of a novel interferometric
Belmonte, C., Garcia-Hirschfeld, J., & Gallar, J. 951–958. method to investigate the relation between lipid
(1997). Neurobiology of ocular pain. Prog. Retin. Funderburgh, J. L., Funderburgh, M. L., Mann, M. layer thickness and tear film thinning. Invest.
Eye Res., 16, 117–156. M., et al. (1991). Physical and biological proper- Ophthalmol. Vis. Sci., 51, 2418–2423.
Berman, E. R. (1981). Biochemistry of the Eye. New ties of keratan sulphate proteoglycan. Biochem. Klyce, S. D. (1981). Stromal lactate accumulation
York: Plenum Press. Soc. Trans., 19, 871–876. can account for corneal oedema osmotically fol-
Bonanno, J. A. (2012). Molecular mechanisms un- Gipson, I. K., & Inatomi, T. (1995). Extracellular lowing epithelial hypoxia in the rabbit. J. Physi-
derlying the corneal endothelial pump. Exp. Eye matrix and growth factors in corneal wound ol., 321, 49–64.
Res., 95, 2–7. healing. Curr. Opin. Ophthalmol., 6, 3–10. Klyce, S. D., & Beuerman, R. W. (1998). Structure
Bron, A. J. (1986). Lacrimal streams: the demon- Gipson, I. K., & Inatomi, T. (1997). Mucin genes and function of the cornea. In H. E. Kaufman, B.
stration of human lacrimal fluid secretion and expressed by the ocular surface epithelium. Prog. A. Barron, & M. B. McDonald (Eds.), The Cor-
the lacrimal ductules. Br. J. Ophthalmol., 70, Retin. Eye Res., 16, 81–98. nea (2nd ed.) (pp. 3–50). Boston: Butterworth-
241–245. Gipson, I. K., Spurr-Michaud, S. J., & Tisdale, A. Heinemann.
Bron, A. J., Benjamin, L., & Snibson, G. R. (1991). S. (1987). Anchoring fibrils form a complex net- Klyce, S. D., Maeda, N., & Byrd, T. J. (1998). Cor-
Meibomian gland disease. Classification and work in human and rabbit cornea. Invest. Oph- neal topography. In H. E. Kaufman, B. A. Bar-
grading of lid changes. Eye, 5, 395–411. thalmol. Vis. Sci., 28, 212–220. ron, & M. B. McDonald (Eds.), The Cornea (2nd
Bron, A. J., Tripathi, R. C., & Tripathi, B. (1997). Glasgow, B. J., Marshall, G., Gasymov, O. K., et al. ed.) (pp. 1055–1075). Boston: Butterworth-
Wolff ’s Anatomy of the Eye and Orbit (8th ed.). (2000). Tear lipocalins: potential scavengers for Heinemann.
London: Chapman and Hall. the corneal surface. Invest. Ophthalmol. Vis. Sci., Knop, E., & Knop, N. (2005). The role of eye-
Bron, A. J., Argüeso, P., Irkec, M., et  al. (2015). 40, 3100–3107. associated lymphoid tissue in corneal immune
Clinical staining of the ocular surface: mecha- Gray, H., Bannister, L. H., Berry, M. M., et  al. protection. J. Anat., 206, 271–285.
nisms and interpretations. Prog. Retin. Eye Res., (1995). Gray’s Anatomy: The Anatomical Basis Knop, E., Knop, N., Zhivov, A., et al. (2011a). The
44, 36–61. of Medicine and Surgery (38th ed.). New York: lid wiper and muco-cutaneous junction anatomy
Butovich, I. A. (2013). Tear film lipids. Exp. Eye Churchill Livingstone. of the human eyelid margins: an in vivo confo-
Res., 117, 4–27. Guillon, J. P. (1998). Non-invasive Tearscope Plus cal and histological study. J. Anat., 218, 449–461.
Corfield, A. P., Carrington, S. D., Hicks, S. J., et al. routine for contact lens fitting. Cont. Lens Ante- Knop, E., Knop, N., Millar, T., et  al. (2011b). The
(1997). Ocular mucins: purification, metabolism rior Eye, 21, S31–S40. international workshop on meibomian gland
and functions. Prog. Retin. Eye Res., 16, 627–656. Hassell, J. R., & Birk, D. E. (2010). The molecular basis dysfunction: report of the subcommittee on
Craig, J. P., & Tomlinson, A. (1997). Importance of of corneal transparency. Exp. Eye Res, 91, 326–335. anatomy, physiology, and pathophysiology of the
the lipid layer in human tear film stability and Hodges, R. R., & Dartt, D. A. (2013). Tear film mu- meibomian gland. Invest. Ophthalmol. Vis. Sci.,
evaporation. Optom. Vis. Sci., 74, 8–13. cins: front line defenders of the ocular surface; 52, 1938–1978.
Dartt, D. A. (1992). Physiology of tear produc- comparison with airway and gastrointestinal Knop, N., Korb, D. R., Blackie, C. A., et al. (2012).
tion (Ch. 10, pp. 65–99). In M. A. Lemp, & R. tract mucins. Exp. Eye Res., 117, 62–78. The lid wiper contains goblet cells and goblet cell
Marquardt (Eds.), The Dry Eye: a Comprehensive Hodson, S. A. (1997). Corneal stromal swelling. crypts for ocular surface lubrication during the
Guide. Berlin: Springer-Verlag. Prog. Retin. Eye Res., 16, 99–116. blink. Cornea, 31, 668–679.
Dartt, D. A. (2009). Neural regulation of lacrimal Hodson, S. A., & Miller, F. (1976). The bicarbonate ion Kronfeld, P. C., McHugh, S. L., & Polyak, S. L.
gland secretory processes: relevance in dry eye pump in the endothelium which regulates the hy- (1943). The Human Eye in Anatomical Transpar-
disease. Prog. Retin. Eye Res., 28, 155–177. dration of the rabbit cornea. J. Physiol., 263, 563–577. encies. Rochester: Bausch & Lomb.
Dilly, P. N. (1985). On the nature and the role of the Hodson, S. A., Guggenheim, J., Kaila, D., et  al. Lander, T., Wirtschafter, J. D., & McLoon, L. K.
sub-surface vesicles in the outer epithelial cells of (1991). Anion pumps in ocular tissues. Biochem. (1996). Orbicularis oculi muscle fibres are rela-
the conjunctiva. Br. J. Ophthalmol., 69, 477–481. Soc. Trans., 19, 849–852. tively short and heterogeneous in length. Invest.
Doane, M. G. (1981). Blinking and the mechanics Hogan, M. J., Alvarado, J. A., & Weddell, J. E. Ophthalmol. Vis. Sci., 37, 1732–1739.
of the lacrimal drainage system. Ophthalmology, (1971). Histology of the Human Eye. Philadel- Lavker, R. M., Tseng, S. C., & Sun, T. T. (2004). Cor-
88, 844–851. phia: Saunders. neal epithelial stem cells at the limbus: looking at
Doughty, M. J., & Zaman, M. L. (2000). Human Hollingsworth, J., Perez-Gomez, I., Mutalib, H. A., some old problems from a new angle. Exp. Eye.
corneal thickness and its impact on intraocular et al. (2001). A population study of the normal Res., 78(3), 433–446.
pressure measures: a review and meta-analysis cornea using an in  vivo, slit-scanning confocal Lawrenson, J. G. (1997). Corneal sensitivity in
approach. Surv. Ophthalmol., 44, 367–408. microscope. Optom. Vis. Sci., 78, 706–711. health and disease. Ophthalmol. Physiol. Opt.,
Dua, H. S., & Azuara-Blanco, A. (2000). Limbal Holly, F. J., & Lemp, M. A. (1977). Tear physiology 17, S17–S22.
stem cells of the corneal epithelium. Surv. Oph- and dry eye. Surv. Ophthalmol., 22, 69–87. Lawrenson, J. G., & Ruskell, G. L. (1991). The struc-
thalmol., 44, 415–425. Jester, J. V., Murphy, C. J., Winkler, M., et al. (2013). ture of corpuscular nerve endings in the limbal
Dua, H. S., Faraj, L. A., Said, D. G., et  al. (2013). Lessons in corneal structure and mechanics to conjunctiva of the human eye. J. Anat., 177,
Human corneal anatomy redefined: a novel pre- guide the corneal surgeon. Ophthalmology, 120, 75–84.
Descemet’s layer (Dua’s layer). Ophthalmology, 1715–1717. Lawrenson, J. G., Reid, A. R., & Allt, G. (1998).
120, 1778–1785. Jones, L. T. (1961). An anatomical approach to Corneal glycoconjugates: an ultrastructural
Efron, N. (1997). Contact lenses and corneal physi- problems of the eyelids and lacrimal apparatus. lectin-gold study. Histochem. J., 30, 51–60.
ology. Biol. Sci. Rev., 9, 29–31. Arch. Ophthalmol., 66, 111–124. Maitchouk, D. Y., Beuerman, R. W., Ohta, T., et al.
Efron, N., & Carney, L. G. (1979). Oxygen levels be- Jones, L. T. (1966). The lacrimal tear system and its (2000). Tear production after unilateral removal
neath the closed eyelid. Invest. Ophthalmol. Vis. treatment. Am. J. Ophthalmol., 62, 47–60. of the main lacrimal gland in squirrel monkeys.
Sci., 18, 93–95. Jordan, A., & Baum, J. (1980). Basic tear flow: does Arch. Ophthalmol., 118, 246–252.
Efron, N., Perez-Gomez, I., Mutalib, H. A., et  al. it exist? Ophthalmology, 87, 920–930. Mantelli, F., & Argüesco, P. (2008). Functions of oc-
(2001). Confocal microscopy of the normal hu- Kessing, S. V. (1968). Mucus gland system of the ular surface mucins in health and disease. Curr.
man cornea. Cont. Lens Anterior Eye, 24, 16–24 conjunctiva: a quantitative normal anatomical Opin. Allergy Clin. Immunol., 8, 477–483.
Erratum in: Cont. Lens Anterior Eye, 24, 83–85. study. Acta Ophthalmol, 95(Suppl.), 1–333. Marfurt, C. F., & Ellis, L. C. (1993). Immunohisto-
Ehlers, N. (1965). On the size of the conjunctival Khoramnia, R., Rabsilber, T. M., & Auffarth, G. U. chemical localisation of tyrosine hydroxylase in
sac. Acta Ophthalmol., 43, 205–210. (2007). Central and peripheral pachymetry mea- corneal nerves. J. Comp. Neurol., 336, 527–531.
Feng, Y., & Simpson, T. L. (2008). Corneal, limbal, surements according to age using the Pentacam Maurice, D. M. (1957). The structure and transpar-
and conjunctival epithelial thickness from opti- rotating Scheimpflug camera. J. Cataract Refract ency of the cornea. J. Physiol., 136, 263–286.
cal coherence tomography. Optom. Vis. Sci., 85, Surg., 33, 830–836. Maurice, D. M. (1984). The cornea and sclera. In
880–883. King-Smith, P. E., Fink, B. A., Fogt, N., et al. (2000). H. Davson (Ed.), The Eye. (3rd ed.) (Vol. 1B) (pp.
Freeman, R. D. (1972). Oxygen consumption by The thickness of the preocular tear film: evidence 1–158). Orlando: Academic Press.
the component layers of the cornea. J. Physiol., from reflection spectra. Invest. Ophthalmol. Vis. McClellan, K. A. (1997). Mucosal defence of the
225, 15–32. Sci., 41, 3348–3359. outer eye. Surv. Ophthalmol., 42, 233–346.
27.e1
27.e2 REFERENCES

McGill, J. I., Holgate, S. T., Church, M. K., et  al. Prydal, J. J., Franc, F., Dilly, P. N., et al. (1998). Kerato- Stiemke, M. M., Watsky, M. A., Kangas, T. A., et al.
(1998). Allergic eye disease mechanisms. Br. J. cyte density and size in conscious humans by digital (1995). The establishment and maintenance of
Ophthalmol., 82, 1203–1214. image analysis of confocal images. Eye, 12, 337–342. corneal transparency. Prog. Retin. Eye Res., 14,
McKenzie, R. W., Jumblatt, J. E., & Jumblatt, M. M. Quantock, A. J., & Young, R. D. (2008). Development 109–140.
(2000). Quantification of MUC2 and MUC5AC of the corneal stroma, and the collagen–proteo- Stoeckelhuber, M., Stoeckelhuber, B. M., & Welsch,
transcripts in the human conjunctiva. Invest. glycan associations that help define its structure U. (2003). Human glands of Moll: histochemi-
Ophthalmol. Vis. Sci., 41, 703–708. and function. Dev. Dyn., 237, 2607–2621. cal and ultrastructural characterization of the
Meek, K. M., & Boote, C. (2009). The use of X-ray Riley, M. V. (1969). Glucose and oxygen utiliza- glands of Moll in the human eyelid. J. Invest.
scattering techniques to quantify the orientation tion by the rabbit cornea. Exp. Eye Res., 8, Dermatol., 121, 28–36.
and distribution of collagen in the corneal stro- 193–200. Sullivan, D. A., & Sato, E. H. (1994). Immunology
ma. Prog. Retin. Eye Res., 28, 369–392. Ruskell, G. L. (1975). Nerve terminals and epithe- of the lacrimal gland. In D. M. Albert, & F. A.
Meek, K. M., & Leonard, D. W. (1993). Ultrastruc- lial cell variety in the human lacrimal gland. Cell Jakobiec (Eds.), Principles and Practice of Oph-
ture of the corneal stroma: a comparative study. Tiss. Res., 158, 121–136. thalmology (pp. 479–486). Philadelphia: W.B.
Biophys. J., 64, 273–280. Ruskell, G. L., & Lawrenson, J. G. (1994). In- Saunders.
Møller-Pederson, T., & Ehlers, N. (1995). A three- nervation of the anterior segment. In M. Sullivan, D. A., Sullivan, B. D., Ullman, M. D., et al.
dimensional study of the human corneal kerato- Guillon, & M. Ruben (Eds.), Contact Lens (2000). Androgen influence on the meibomian
cyte density. Curr. Eye Res., 14, 459–464. Practice (pp. 225–237). London: Chapman gland. Invest. Ophthalmol. Vis. Sci., 41, 3732–3742.
Nishida, T., & Tanaka, T. (1996). Extracellular ma- and Hall. Takahashi, Y., Watanabe, A., Matsuda, H., et  al.
trix and growth factors in corneal wound heal- Sack, R. A., Tan, K. O., & Tan, A. (1992). Diurnal (2013). Anatomy of secretory glands in the eye-
ing. Curr. Opin. Ophthalmol., 7, 2–11. tear cycle: evidence for a nocturnal inflamma- lid and conjunctiva: a photographic review. Oph-
Oliveira-Soto, L., & Efron, N. (2001). Morphology tory constitutive tear fluid. Invest. Ophthalmol. thal. Plast. Reconstr. Surg., 29(3), 215–219.
of corneal nerves using confocal microscopy. Vis. Sci., 33, 626–640. Tiffany, J. M. (1991). The viscosity of human tears.
Cornea, 20, 374–384. Saonanon, P. (2014). Update on Asian eyelid anat- Int. Ophthalmol. Clin., 15, 371–376.
Paulsen, F. P., Thale, A. B., Hallmann, U. J., et al. omy and clinical relevance. Curr. Opin. Ophthal- Tiffany, J. M. (1995). Physiological functions of the
(2000). The cavernous body of the human ef- mol., 25, 436–442. meibomian glands. Prog. Retin. Eye Res., 14, 47–74.
ferent tear ducts: function in tear outflow Scott, J. E. (1991). Proteoglycan: collagen interac- Tiffany, J. M. (1997). Tears and conjunctiva. In
mechanism. Invest. Ophthalmol. Vis. Sci., 41, tions and corneal ultrastructure. Biochem. Soc. J. J. Harding (Ed.), Biochemistry of the Eye
965–970. Trans., 19, 877–881. (pp. 1–15). London: Chapman and Hall.
Pfister, R. R. (1973). The normal surface of the cor- Seifert, P., Spitznas, M., Koch, F., et al. (1993). The Tsubota, K. (1998). Tear dynamics and dry eye.
neal epithelium: a scanning electron microscopic architecture of human accessory lacrimal glands. Prog. Retin. Eye. Res., 17, 565–596.
study. Invest. Ophthalmol. Vis. Sci., 12, 654–668. Ger. J. Ophthalmol., 2, 444–454. Wilson, S. E. (2000). Role of apoptosis in wound
Pflugfelder, S. C. (1998). Tear fluid influence on the Shaheen, B. S., Bakir, M., & Jain, S. (2014). Corneal healing in the cornea. Cornea, 19, S7–S12.
ocular surface. Adv. Exp. Med. Biol., 438, 611–617. nerves in health and disease. Surv. Ophthalmol., Wilson, S. E., & Hong, J. W. (2000). Bowman’s layer
Pflugfelder, S. C., Liu, Z., Monroy, D., et  al. 59, 263–285. structure and function. Critical or dispensable
(2000). Detection of sialomucin complex Sherrard, E., Novakovic, P., & Speedwell, L. (1987). to corneal function? A hypothesis. Cornea, 19,
(MUC4) in human ocular surface epithelium Age related changes of the corneal endothelium 417–420.
and tear fluid. Invest. Ophthalmol. Vis. Sci., 41, and stroma as seen in vivo with the specular mi- Wirtschafter, J. D., Ketcham, J. M., Weinstock, R. J.,
1316–1326. croscope. Eye, 1, 197–203. et al. (1999). Mucocutaneous junction as a ma-
Prydal, J. J., Artal, P., Woon, H., et al. (1992). Study Srinivasan, S., Menzies, K., Sorbara, L., et al. (2012). jor source of replacement palpebral conjunctival
of human preocular tear film thickness and Infrared imaging of meibomian gland structure epithelial cells. Invest. Ophthalmol. Vis. Sci., 40,
structure using interferometry. Invest. Ophthal- using a novel keratograph. Optom. Vis. Sci., 89, 3138–3146.
mol. Vis. Sci., 33, 2006–2011. 788–794.

You might also like