Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Jul 7.
Published in final edited form as: Ocul Surf. 2017 Jul 20;15(3):366–403. doi: 10.1016/j.jtos.2017.03.006

TFOS DEWS II Tear Film Report

Mark DP Willcox a,1,*, Pablo Argüeso b, Georgi A Georgiev c, Juha M Holopainen d, Gordon W Laurie e, Tom J Millar f, Eric B Papas a, Jannick P Rolland g, Tannin A Schmidt h, Ulrike Stahl i, Tatiana Suarez j, Lakshman N Subbaraman i, Omür O Uçakhan k, Lyndon Jones i
PMCID: PMC6035753  NIHMSID: NIHMS978473  PMID: 28736338

Abstract

The members of the Tear Film Subcommittee reviewed the role of the tear film in dry eye disease (DED). The Subcommittee reviewed biophysical and biochemical aspects of tears and how these change in DED. Clinically, DED is characterized by loss of tear volume, more rapid breakup of the tear film and increased evaporation of tears from the ocular surface. The tear film is composed of many substances including lipids, proteins, mucins and electrolytes. All of these contribute to the integrity of the tear film but exactly how they interact is still an area of active research. Tear film osmolarity increases in DED. Changes to other components such as proteins and mucins can be used as biomarkers for DED. The Subcommittee recommended areas for future research to advance our understanding of the tear film and how this changes with DED. The final report was written after review by all Subcommittee members and the entire TFOS DEWS II membership.

Keywords: Dry eye disease, Evaporation, Lipidome, Mucin, Osmolarity, Proteome, Tear film, Tear film stability, Tears

1. Overview of the tear film in health

A stable preocular tear film is a hallmark of ocular health, largely because it forms the primary refracting surface for light entering the visual system and it protects and moisturizes the cornea. The three layered model of the tear film proposed by Wolff [1,2] has had an overwhelming allure because it is simple and logical: a mucin layer covering the ocular surface and lowering the supposed hydrophobicity of the epithelial cells; an aqueous layer to nurse the exposed ocular epithelium by providing lubricity, some nutrients, antimicrobial proteins and appropriate osmolarity; and a lipid layer to prevent loss of the aqueous layer through overspill and evaporation. There is a continual return to this three layered model, despite Doane stating over 20 years ago that the three layered structure “is a considerable simplification of reality” [3]. This has generally limited novel perspectives that might lead to a clearer understanding of the dynamics, structure and function of the tear film and the changes that occur to cause dry eye. The precorneal tear film behaves as a single dynamic functional unit [4] with different compartments.

Laxity of terminology means that there is ready acceptance of information that may not be entirely correct. For instance, “tear osmolarity is approximately 302 mOsm/L” is often accepted terminology, but in reality such a value is for tears sampled from within the lower tear meniscus. While it may represent the osmolarity of the tear film spread over the ocular surface, there is no evidence of this. A consequence of being aware of where the samples being measured are coming from and how they are collected may lead to a more cautious approach to extrapolating data to the tear film that covers the ocular surface and, ultimately, a better understanding of its composition, structure and spatial distribution. Optical coherence tomography (OCT) has allowed non-invasive measurements of both the upper and lower menisci in terms of height, area, and curvature of the surface and while the upper and lower meniscus in an individual appear to be identical in these parameters, none of these parameters correspond to central tear film thickness [5], but lower meniscus height seems to correspond to the volume of mucoaqueous tears [6,7].

When the eyes are open the tears are distributed in three compartments, which are the fornical compartment (which occupies the fornix and retrotarsal space), the tear menisci and the preocular tear film. The fornical compartment is assumed to be narrowest in the region of the lid wiper of the lid margin, which is directly apposed to the globe. The preocular tear film overlies the exposed conjunctiva and cornea [8]. The precorneal tear film follows the contours of the cornea, and is usually highly stable [9]. The pre-bulbar film follows the varying contours of the bulbar conjunctiva. The preocular tear film is the whole tear component that is spread over the exposed surface of the eye. Results from studies using ultrahigh resolution OCT has resolved the debate over the thickness of the tear film. It is extraordinarily thin, 2–5.5 μm thick over the corneal region (precorneal tear film), and these data concur with estimates of tear film thickness using interferometry techniques [810]. The tear film is so thin that the roughness of the corneal surface (~0.5 μm) cannot be ignored [11]. Neither the tear film thickness over the conjunctival region, nor the roughness of the conjunctival surface has been measured. Water has a high surface tension and therefore to form such a thin film of water without it collapsing onto the surface or forming lenses, the surface on which it spreads has to have similar properties to water and the surface tension of the water at the air interface has to be lowered [12,13].

The apical surfaces of the corneal and conjunctival epithelia have transmembrane mucins [14], which increase the adhesion tension for water, facilitating the spread of the tears across the ocular surface. Transmembrane mucins attached to the microplicae of the epithelial cells extend up to 500 nm (0.5 μm) into the tear film [15,16]. They also constitute a line of defense for the epithelial cells against infection and injury [17,18].

Much remains to be learned about the mucoaqueous component of the preocular tear film and whether it is the same within all compartments. In addition to oxygen, metabolites and electrolytes, the tear film contains antimicrobial peptides, proteins and soluble immunoglobulins that protect the ocular surface from infection. The sensitivity of modern proteomics techniques has allowed the identification of more than 1500 proteins [19], and more than 200 peptides originating from several of those proteins [20]. The nature of the vast majority of these proteins and peptides reflects that tears are also a mechanism for removal of cellular debris that occurs due to the turnover of ocular epithelial cells. In addition, sensitive lipid studies also show that tears contain a lipid profile similar in ratio to meibomian lipids, but with a relative abundance of phospholipids [21].

Shortly after a blink, the mucoaqueous component of the pre-ocular tear film is believed to become physically isolated from the upper and lower menisci, such that diffusion between these compartments does not occur [22,23]. This isolation has been observed as a black line at the ocular margins in fluorescein-labeled tears.

The tear film lipid layer is approximately 40 nm thick [24], lowers the surface tension at the air interface of the preocular tear film and results in spreading of the tear film over the ocular surface. A feature of the preocular tear film is that it resists evaporation, and it is purported that the tear film lipid layer is responsible for this [2527,67]. The complete nature of the tear film lipid layer is unknown, but it is likely it has surfactant molecules at the mucoaqueous interface and lipophilic molecules at the air interface. Unlike the aqueous component of the preocular tear film, which appears to be isolated shortly after a blink, the lipid layer of the tear film appears to be continuous over the menisci and indeed continues to move upwards over the ocular surface following a blink. Observations using various interference techniques at different magnifications show that the lipid layer is variable in thickness across the ocular surface. This movement and continuum from the meibomian gland orifices, and direct observation of secretions from the meibomian glands onto the ocular surface, indicate that the tear film lipid layer is almost entirely derived from meibomian gland secretions. It is unknown if lipids from the tear film lipid layer move into the mucoaqueous compartment of the preocular tear film, or if lipids from other ocular tissues (origin unknown) transverse the mucoaqueous compartment and adsorb to the tear film lipid layer. By examining the shear rheology of films of meibomian lipids in vitro and comparing them with other lipids, there is evidence that meibomian lipid films spread over a mucoaqueous subphase, causing the subphase to resist collapse as it thins [28]. Dilatational rheology studies also confirmed that the viscoelastic films of tear lipids, meibum and contact lens lipid extracts are predominantly elastic, which may enhance their capability to stabilize the air/tear surface [2931].

2. Biophysical measurements of the tear film

2.1. Tear film structure and dynamics

There is evidence that, in the above described three layer structure of the tear film, the mucin layer has a decreasing gradient of concentration from the epithelium towards the aqueous layer [32]. It is also commonly considered that the aqueous and mucin layers are a single layer of mucoaqueous gel (referred to hereafter as the mucoaqueous layer) [33].

The tear film lipid layer is derived from meibum secreted from the lid margins and is spread onto the tear film with each blink, driven by surface tension forces. It plays an important role in stabilizing the tear film and in the past has been thought to play a key role in retarding tear evaporation [2527,67]. The lipid layer can be investigated with interferometry techniques. The color and brightness of the interference images are analyzed to yield lipid layer thickness [24,3437]. The thickness of the lipid layer has been reported to be from 15 to 157 nm, with a mean of 42 nm [24].

Evidence from reflection spectra of the precorneal tear film suggested the tear film has a thickness of approximately 2 μm [9]. OCT techniques find the thickness of the tear film to range from 2 to 5.5 μm [10,3842].

To elucidate the structure of the tear film, studies have made use of multiple methods. These include combining a wavefront sensor with OCT [43], using fluorescein tear breakup time (TBUT) and Schirmer test [43], applying fluorescein and assessing using a rotating Scheimpflug camera (Pentacam, Oculus, Germany) [44], simultaneously recording videos of fluorescence and imaging tear film lipid layer [45], and using dual thermal-fluorescent imaging [46].

The bulk of the tear volume and flow is via secretion from the lacrimal gland [47,48], with a smaller portion secreted by the conjunctiva [49]. In animal studies, tears can be produced by the accessory lacrimal glands in the conjunctiva even after removal of the main gland [50]. Both parasympathetic and sympathetic nerves innervate the main lacrimal glands [51,52] and a few sensory nerves have also been identified [51]. The nerves are located in close proximity to acinar, ductal, and myoepithelial cells as well as being close to blood vessels [51,52]. Stimulation of the lacrimal gland and secretion occur via the cornea –trigeminal nerve–brain-stem–facial nerve–lacrimal gland reflex arc. Afferent sensory nerves of the cornea and conjunctiva are activated by stimulation of the ocular surface. Efferent parasympathetic and sympathetic nerves are then activated to stimulate secretion from acinar and tubular cells in the lacrimal gland [53].

Tears have been classified into four broad types - basal, reflex, emotional and closed-eye. (see review by Craig et al.) [54]. Basal tears (sometimes referred to as open-eye tears) are tears that constitutively coat the eye and are deficient in dry eye. Reflex tears are produced upon stimulation of the ocular surface (for example by onion vapor) or stimulation of the reflex arc (for example by nasal stimulation of the sneeze reflex). Emotional tears are also produced upon stimulation, but in this case via emotions such as sadness. Closed-eye tears are those that can be collected from the ocular surface immediately after a period of sleep. Basal, reflex and emotional tears are produced mainly from the lacrimal glands via the neural arc [55], but differ in their constitution, for example, the concentration of various proteins change [54]. Secretion from the lacrimal gland is greatly reduced during sleep, and so the constitution of closed-eye tears is somewhat different to that of other types with, for example, an increased amount of serum-derived proteins leaking from the conjunctival blood vessels [54].

A two-step process of tear film deposition through a blink has been proposed [56]. In the first step, the upper lid pulls a layer of tears over the cornea by capillary action; in the second step, the lipid layer drifts upward, which may drag up aqueous tears along with it. The upward drift of the lipid layer can be observed using interferometry imaging approaches [57]. After the blink, tear film redistribution occurs due to the negative hydrostatic pressure within the nascent menisci. This draws liquid from the forming tear film and eventually causes the precorneal portion to separate from the menisci. The boundary can be observed as a black line of reduced fluorescence in the fluorescein-stained tear film, indicating where the aqueous layer is thin but the lipid layer remains intact [58,59].

Tears flow from the supply region towards the puncta, located on the lids near the nasal canthi, to facilitate their turnover and removal [60,61]. Tear turnover rate has been estimated to be 16 ± 5%/min [6264]. Between blinks, thinning of the tear film occurs, which can be observed using several different approaches [40,57,65,66]. Most of the observed tear thinning between blinks is due to evaporation [2527,57,65,67,68].

Tear production, turnover and volume can be estimated by several methods, but there is limited correlation between different tests [69]. Accordingly, a combination of tests should provide a more reliable diagnosis and increase the specificity and sensitivity of dry eye diagnosis [70]. The phenol red thread test (Hamano test) [71] is a measurement of tear volume or change in tear volume with time, by observation of the amount of wetting of a phenol red dye impregnated cotton thread placed over the inferior eyelid. The Schirmer test [72] is a measure of tear production and is undertaken by observing the wetting of a standardized paper strip. Historically, the Schirmer I test is performed without anesthesia and thus measures predominantly reflex tearing. A variation on the Schirmer I test involves use of topical anesthesia and claims to reflect the basal secretion of tears, although a contribution from reflex tearing cannot be discounted [626]. Tear volume can also be measured by fluorophotometric assessment, and demonstrates an apparent normal human volume of approximately 8 ± 3 μl [47,62]. Tear meniscus height (TMH) is linearly proportional to the lacrimal secretory rate [47]. Differences in TMH and radius of curvature can be used to aid diagnosis of dry eye [7,73,74]. Tear clearance rate is the rate at which the preocular tear film or an instilled marker of the tears is removed from the tear film by dilution or drainage from the tear volume [75]. Tear clearance rate measurement is seldom performed in the clinical setting. Tear dynamics can be estimated by dividing the value of the Schirmer test with anesthesia by the tear clearance rate, giving the Tear Function Index [76]. This value has been shown to have greater sensitivity for detecting dry eye than either one of these tests alone [76].

2.2. Tear film stability on eye and ocular surface wettability

A stable precorneal tear film has long been viewed as one of the hallmarks of ocular health, largely because it provides the primary refracting surface for light entering the visual system as well as creating a protective and lubricated environment for the tissues of the palpebral and bulbar surfaces. Unlike some other species, whose tears can remain stable for many minutes [77], the human tear film tends to collapse or “break up” in under half a minute or so, unless it is re-established by the act of blinking. While all individuals will manifest this behavior if blinking is prevented for long enough, rapid appearance of regions of localized drying is viewed as evidence for tear film disorder, particularly in dry eye, and so observations of stability are commonly and frequently performed as a diagnostic aid.

Experiments have shown that tear film thinning and breakup occur mainly as a result of evaporation from the tear film, rather than due to fluid flow, whether that be tangentially within the film itself, or radially across the ocular surface [78,79]. Using reflectivity and tear fluorescein as respective indicators for lipid and mucoaqueous layer thickness, King-Smith et al. suggested a lack of correspondence between dry eye and both lipid layer thickness and thinning rate [45]. Further, thinning rate was not affected by apparent thickening of the lipid layer with lipid emulsion-based eye drops [24], suggesting that the lipid was a poor barrier to evaporation [45]. Perhaps it is the whole healthy preocular tear film that resists evaporation from the ocular surface, and hence thinning? Measurement of TBUT may provide a better indicator of the ability of the preocular tear film to prevent evaporative losses.

Acquiring a TBUT is a relatively simple task, but interpreting the result is not straightforward because of its inherent variability [80,81]. A number of approaches to improving repeatability have been suggested, including taking multiple readings and averaging or selecting a subset of values [82,83], minimizing the amount of fluorescein instilled [8486] and, most significantly, eliminating the use of fluorescein altogether. This last approach, via a number of different methods, provides a non-invasive breakup time (NIBUT) value. Tear film additives are avoided, the examination environment should ideally introduce no additional sources of heat, air movement, humidity etc., and head posture and blinking behavior are standardized. Inevitably however, the extent to which these conditions are achieved varies somewhat between methods.

Early efforts in acquiring a NIBUT projected a grid pattern [87,88] or keratometry mires [89,90] onto the surface of the tear film and viewed their distortion in time after a blink. While it is still possible to use keratometer-based methods in clinical situations, a degree of subjectivity is involved in judging when the image distortion first occurs. Increasing sophistication in both image capture and computational capability has led to a refinement of the technique, using both automated detection and more detailed targets. In most cases, these targets are identical to those originally developed and used for measuring corneal shape (keratoscopes) and consist of multiple concentric rings whose angular subtense is sufficient to cover more or less all of the visible cornea. The image of this target is reflected from the anterior surface of the tear film and captured for subsequent analysis [9198]. Typically, multiple, sequential images are acquired during the inter-blink period and image analysis software utilized to automatically detect the onset of areas of breakup.

Although repeatability has not been reported for all the devices using this approach, the available data are in reasonable agreement that they operate with a coefficient of variation of around 10% [92,94], which is roughly three times better than traditional TBUT measurement [92]. Despite this improvement, the range of values reported for normal individuals is broad, being from about 4 to 19 s (Table 1). It may be that this reflects the different algorithms being used to extract breakup data among the various instruments, a suspicion that is strengthened by the observation that the corresponding dry eye breakup times are consistently about half those given for normal eyes. Thus, while inter-instrument comparisons are likely to be difficult to interpret, data derived from a given instrument type appear reasonably reliable and offer quite good sensitivity and specificity in distinguishing dry-eyed individuals from normals (Table 1). Further details may be obtained from the Tear Film and Ocular Surface Society’s Dry Eye Workshop II (TFOS DEWS II) Diagnostic Methodology report [99].

Table 1.

Summary of non-invasive breakup time (NIBUT) measurements in normal and dry-eyed subjects, together with discrimination diagnostic metrics. NIBUT = non-invasive breakup time; AUC = area under the curve of a receiver operating characteristic graph plotting sensitivity vs. 1-specificity.

Author (Reference) NIBUT Normal (sec) NIBUT Dry Eye (sec) AUC Sensitivity Specificity Instrument Principle
Hong et al. 2013[94] 4.3 ± 0.3
n = 41
2.0 ± 0.2
0.83 84.1 75.6 Oculus keratograph Videokeratoscopy
Gumus et al. 2011[97] 4.9 ± 1.6
n = 25
2.4 ± 2.5; Mild n = 23
1.2 ± 1.8; Moderate n = 11
0.4 ± 0.5; Severe n = 11
82.2 88.0 Tomey RT7000 Videokeratoscopy
Downie 2015 [92] 19.4 ± 5.3
n = 17
7.9 ± 4.9
n = 28
0.92 81.5 94.4 Medmont E300 Videokeratoscopy
Koh et al. 2016[95] 9.7 ± 6.7
n = 31
4.6 ± 1.3
n = 49
Keratograph M5 Videokeratoscopy
Su et al. 2016[109] 4.5 ± 0.9
n = 31
2.1 ± 1.1
n = 42
0.88 80.0 89.0 IT-85, United Integrated Services Co Thermography

Ocular surface thermography has been used to measure NIBUT, on the principle that breakup is associated with evaporative cooling and therefore thinning areas in the tear film show up as cool spots in the thermograph [46]. The technique is suggested to be reliable [100], although test-retest confidence intervals have not been made available so far. While data are limited to a single study, NIBUT derived from this method appear similar to those from the lower end of the videokeratoscopy range. Again, dry-eyed subjects yield breakup times that are about half those of normals, with levels of sensitivity and specificity being similar to those from video-keratoscopy (Table 1).

All the systems discussed so far are commercially available and so could feasibly be used in routine clinical practice. The following discussion deals with instruments that are more complex and/or unlikely to be applicable outside a research setting. Recently, lateral shearing interferometry has been used to monitor changes in tear film stability. This instrument uses an optical wedge to laterally shift the wavefront reflected from the tear film surface and rotate it so that it can be made to interfere with itself [101,102]. Information about the shape of the reflecting surface is contained in this wavefront and can be extracted from the resulting interference pattern. Note that this differs from colored fringe methods, such those of Guillon [103], or the interferometer developed by Doane [104], both of which rely on interference between light reflected from different surfaces within the tear film, such as the front and back of the lipid layer. Using fast Fourier transformation, images derived with the shearing technique can be processed to generate a surface stability index parameter (M2). Sequential image acquisition allows M2 to be followed over the blink cycle, in real time, at a resolution determined by the video frame rate. It is claimed that this method is relatively insensitive to eye movements and the degree of dryness of the surface being measured [105] and is better able to discriminate dry eyed subjects from normal than either dynamic area high speed videokeratoscopy or wavefront sensing [106].

Another technique that may be developed for clinical application is the use of double pass methods. The basis for double pass methods is that the view of the retina obtained using a double pass optical system, in which the image forming light traverses all the optical surfaces of the eye twice (once on entry and again on exit), will be affected by scattering from all these surfaces, including the tear film. Thus, analysis of double-pass retinal images on the time scale of the blink cycle may provide an indirect measure of tear film stability. Deteriorations in image quality metrics such as intensity distribution index [107], Strehl ratio, modulation transfer function cut-off frequency and objective scattering index [108] are observed in dry eye. More data are needed to establish the diagnostic ability of this approach for discriminating dry eye.

2.3. Vision quality

An association between dry eye and compromised visual acuity postulated by Rieger [110] is poorly documented using high contrast letter acuity [111]. However, based on the measurement of “functional visual acuity”, whereby acuity is measured after suspension of blinking for several seconds, dry eyed subjects do significantly worse than normals [111]. Delayed blinking generates subtle wavefront aberrations that more rapidly give rise to higher order aberrations after the blink in dry eye individuals [112]. In an effort to directly link tear film changes to visual loss, a three channel optical system that allowed concurrent measurement of letter contrast acuity, TBUT and refractive aberrations was constructed. Although the data reported were for contact lens wearing eyes only, progression of TBUT was clearly associated with both visual performance reduction and declining optical quality [113].

2.4. Tear osmolarity

Tear film osmolarity has been described as a single measurement that gives insight into the balance between tear production, evaporation, drainage and absorption [114]. In the 1995 National Eye Institute/Industry Workshop [115], tear hyperosmolarity was included as a global feature of dry eye and in the 2007 TFOS DEWS definition and classification report [116] tear film hyperosmolarity was identified as one of the two core mechanisms of dry eye and included in the definition. This section will mainly focus on new knowledge gained since the 2007 reports.

The terms tear film osmolarity and osmolality have frequently been employed interchangeably, with osmolarity being the more common term, and is the term used throughout this manuscript [117,118]. Both refer to the amount of osmotically active particles, although with small differences [119] that are clinically irrelevant. Previously, tear film osmolarity was primarily measured via freezing point depression or vapor pressure osmometry, but a series of factors have limited their usage in clinical settings [118121]. Clinical evaluation of tear osmolarity has increased with the introduction of a new osmometer that collects a 50 nL tear sample and analyzes its electrical impedance (TearLab, San Diego, CA, USA).

Mean tear film osmolarity values in normal participants range from 270 to 315 mOsm/L [122137], with an overall average of 300 mOsm/L, which is similar to the value stated by Tomlinson et al., who reviewed studies between 1978 and 2004 [138]. There seems to be no statistical or clinically relevant effect on tear osmolarity of age [128,139141], race [140,142,143], hormonal fluctuation in women with a regular menstrual cycle [131,144], or oral contraceptive pill use [131,144]. Vehof et al. investigated the influence of genetic factors on dry eye in female twins and found an estimated heritability of 40% for osmolarity [145]. Variation between normal right and left eyes is 6.9 ± 5.9 mOsm/L [146]. Data on the effect of sex on tear osmolarity remain equivocal, with Lemp et al. [128] and Versura et al. [126] reporting no significant effect of sex, which is in agreement with a previous review of the literature [54], whilst Fuerst et al. [140] show significantly higher tear film osmolarity in men (311.8 vs. 302.3 mOsm/L). An increase in tear osmolarity has been observed after sleep deprivation, exposure to high altitude and religious fasting [135,147,148]. In accordance with previous studies, prolonged eye closure has been found to lead to tear hypo-osmolarity but data about diurnal variations are equivocal [130,134,139,142,149]. Some authors [130,142] show a shift towards lower values mid-day, followed by an increase, whilst others [134,139,149] show no significant effect of time. There is a positive relationship between plasma osmolarity and tear osmolarity and both are raised in patients with dry eye disease or with systemic dehydration [150152].

The concentration of electrolytes in the mucoaqueous layer mainly determines the osmolarity of the normal tear film, and secretion mechanisms and contribution to tear film osmolarity have been previously summarized [117,119]. Various mathematical models, as summarized and furthered by Braun et al., have stated that the osmolarity across the ocular surface is different to that measured in the tear meniscus [153]. During the blink interval the tear film thins over the cornea, mainly due to evaporation, leading to a hyperosmotic shift [153,154]. The level of hyperosmotic shift depends on the thinning rate, which is driven by the evaporation rate [154]. In the event of a low thinning rate, such as 1 μm/min, tear film osmolarity over the ocular surface will increase from 300 mOsm/L to 332 mOsm/L over a 25 s period, but will increase to 1830 mOsm/L in the case of a rapid thinning rate of 20 μm/min [153]. During tear film breakup, local spikes of tear osmolarity around 1900 mOsm/L have been predicted [153]. Those rates are significantly higher than those observed in the tear meniscus, partly due to the mixing of the tear fluid from the ocular surface with that in the meniscus during the blink and secretion of new tears. In the non-dry eye, the predicted osmolarity difference between the tear film over the ocular surface and in the tear meniscus is fairly small, but is predicted to increase in dry eye, particularly when there is increased evaporation with reduced TMH [68]. The peak in tear film osmolarity during tear film breakup is supported by studies investigating ocular comfort sensations. A slow increase in ocular discomfort has been demonstrated during tear thinning, with a sharp increase at tear film breakup or before a blink [155,156]. Liu et al. evaluated the subjective response during tear film breakup and tried to match the experienced sensation by instilling hyper-osmolar drops [157]. The detection threshold for NaCl drops was 454 ± 14 mOsm/kg, with overall discomfort increasing as osmolarity increased. On average, a salt solution with 809 mOsm/kg was needed to evoke the same ocular response as during TBUT, ranging from 696 to 972 mOsm/kg [157]. The link between tear evaporation, tear thinning and tear film osmolarity is further supported in clinical studies showing a significant correlation between increased evaporation, tear film osmolarity and decreased tear stability, or automated measures of tear film surface quality breakup time being a clinical marker for tear hyperosmolarity in moderate to severe dry eye [92,158].

As summarized in the 2007 TFOS DEWS report, tear hyper-osmolarity in dry eye has been attributed to increased evaporation or a higher impact of evaporation in low volume [116,159]. Mean tear film osmolarity values for dry eye in studies until 2008 ranged between 311 and 360 mOsm/L, with an average of 326.9 mOsm/L [119,138]. Table 2 summarizes the values obtained in studies since 2009 that aimed to establish osmolarity values in dry eye or investigated the feasibility of osmolarity measurements in the diagnosis or treatment success of dry eye. Using different degrees of dry eye severity and etiologies, mean values range between 297 and 337 mOsm/L, with an overall mean of 315. Higher tear film osmolarity values have been reported with increased dry eye severity grade [126,128,160164]. For example, normal, mild/moderate, and severe dry eyes have average tear osmolarity values of 302 ± 8 mOsm/L, 315 ± 11 mOsm/L and 336 ± 22 mOsm/L, respectively [162]. Suzuki et al. and Sullivan et al. found a significant correlation between dry eye severity scores and measured tear film osmolarity (r = 0.47 and r = 0.74, respectively) [160,162]. Despite similar outcomes, it must be noted that the criteria for dry eye severity were not uniform among studies, with authors using different cutoff values, different calculation of scores and different decisions about the inclusion of tear film osmolarity as a diagnostic factor for dry eye severity, which could have introduced selection bias [116].

Table 2.

Tear film osmolarity values in dry eye and non-dry eye subjects.

Author Testing Method Measurement from one eye, combined eyes, or worst eye Non-Dry Eye (mOsm/L) Dry Eye (mOsm/L) Change % Dry Eye Group Definition
Khanal et al., 2009 [123] Freezing point depression NK 308.39 ± 9.29 330.01 ± 13.34 6.6 Aqueous deficient; TFOS DEWS
325.57 ± 14.76 5.3 Evaporative dry eye; TFOS DEWS
Messmer et al., 2010 [122] a Electrical impedance One eye 307.1 ± 11.3 308.9 ± 14.0 0.6b ≥3 of the following criteria were fulfilled:(1) ocular surface disease index (OSDI) > 15; (2) staining of the cornea in the typical interpalpebral area; (3) staining of the conjunctiva in the typical interpalpebral area; (4) tear film breakup time < 7 s; (5) Schirmer test < 7 mm in 5 min; (6) the presence of blepharitis or meibomitis.
Suzuki et al., 2010 [160] a Freezing point depression One eye 309.7 ± 22.3 Modified TFOS DEWS Grades 1–4
Tomlinson et al., 2010 [125] a Electrical impedance One eye 308 ± 6.2 321 ± 7.2 4.1 NIBUT <10 s, Schirmer I test <5 mm, and positive symptoms
Freezing point depression 310 ± 7.2 323 ± 14.7 4.0
Versura et al., 2010 [126] Electrical impedance Worst eye 295.5 ± 9.8 298.1 ± 10.6 0.9b Mild; OSDI 13–21, meibomian gland dysfunction variably present
306.7 ± 9.5 3.7 Moderate; OSDI 23–32, MGD variably present, TBUT ≤ 10 s, Schirmer ≤ 10 mm
314.4 ± 10.1 6.0 Severe; OSDI 33–100, MGD frequent, TBUT ≤ 5 s, Schirmer ≤ 5 mm, moderate to marked corneal and conjunctival staining
Jacobi et al., 2011 [127] a Electrical impedance One eye 301 (298–304) 320 (301–324) 5.9 TFOS DEWS Grade 3
Lemp et al., 2011 [128] Electrical impedance Worst eye 300.8 ± 7.8 315 ± 10.4 4.5 Mild/Moderate; modified TFOS DEWS classification system
336.7 ± 22.2 10.7 Severe; modified TFOS DEWS classification system
Utine et al., 2011 [164] Electrical impedance Both eyes 294.9 ± 8.3 301.9 ± 11.4 2.3 Primary Sjögren syndrome
Li et al., 2012 [130] Electrical impedance Both eyes 298 ± 14.2 304 ± 10.8 2.0b OSDI >12 and a TBUT of <10 s or Schirmer test score of <5 mm
Sullivan et al., 2012 [161] Electrical impedance Both eyes 307.1 ± 9.9 Mild/Moderate; modified TFOS DEWS classification system
324.3 ± 19.8 Severe; modified TFOS DEWS classification system
Szalai et al., 2012 [129] a Electrical impedance Both eyes 303.5 ± 12.9 296.8 ± 16.5 −2.3b Non-Sjögren syndrome; at least 2 of the following 3 objective dry eye tests: Schirmer I value < 10 mm per 5 min, TBUT <10 s, fluorescein corneal staining >4 bright spots
303.4 ± 17.2 0b Sjögren syndrome
Bunya et al., 2013 [167] Electrical impedance Both eyes 314.5 ± 18.0 Severe Sjögren syndrome
Chen et al., 2013 [131] Electrical impedance Both eyes 301.6 ± 14.2
Hassan et al., 2013 [132] Electrical impedance Both eyes 303.62 ± 12.29
Keech et al., 2013 [133] Electrical impedance Both eyes 304.0 ± 8.4 334.0 ± 25.6 9.0 Symptomatic participants; OSDI ≥20
Amparo et al., 2014 [168] a Electrical impedance Worst eye 312.8 ± 20.3 A previous diagnosis of dry-eye disease
Berchicci et al., 2014 [169] Electrical impedance Both eyes 314 ± 22.1 Graft-Versus-Host Disease
Garcia et al., 2014 [134] Freezing point depression Both eyes 270 ± 4.4
Koktekir et al., 2014 [135] Electrical impedance Both eyes 285.6 ± 8.2
Bunya et al., 2015 [137] Electrical impedance Both eyes 312 (299–326) 307 (296–325) −1.6b Blepharitis
314 (306–321) 0.6 Sjögren syndrome
Downie 2015 [92] Freezing point depression Both eyes 296.2 ± 9.9 325.0 ± 17.8 8.9 OSDI score of ≥18, tear osmolarity ≥ 316 mOsm/L and at least one (of three) consecutive fluorescein TBUT of <10 s, in at least one eye
Na et al., 2015 [170] Electrical impedance Both eyes 296 300 1.0 Graft-Versus-Host Disease
Schargus et al., 2015 [171] Electrical impedance Worst eye 300 ± 16.5 337 ± 36 11.0 Ocular Graft-Versus-Host Disease
Tukenmez-Dikmen et al., 2015 [163] a Electrical impedance Both eyes 308.1 ± 8.5 TFOS DEWS Grade 1&2
318.9 ± 12.8 TFOS DEWS Grade 3&4
a

Do not provide information on the use of artificial tears prior to measurements. All other studies stipulated that no artificial tears were to be used, with washout periods ranging from no use at the visit to 1 month.

b

Results were not statistically different between control and dry-eye groups.

Although there can be large within-subject fluctuations [130,137,141,149], Eperjesi et al. proposed that changes of 33 mOsm/L or higher can be considered clinically relevant [141]. The instability of osmolarity readings has been attributed to compromised homeostasis of the dry eye tear film. Differences between readings can be considered a marker for tear film instability [128,133,146]. Such interpretation may be supported by studies showing higher variability between measurements or eyes with higher osmolarity values or more severe dry eye [128,133,137,139,165].

Despite a general shift in tear film osmolarity with dry eye, there is a large overlap in osmolarity values between normal and dry eye participants. Sensitivity and specificity measurements of osmolarity for dry-eye diagnosis using a threshold of 294 mOsm/L were 67% and 46% respectively [143], 40% and 100% using a threshold of >310 mOsm/L and Sjögren syndrome patients [166]. For a more in-depth analysis of statistical measures of performance and potential selection bias of subjects included, please refer to the TFOS DEWS II Diagnostic Methodology report [99]. Future studies are required to establish consistent cut-off values for dry eye, which are ideally derived from a study population where osmolarity was not a diagnostic criterion and where the sensitivity and specificity are tested in an independent cohort.

2.5. Tear ferning

Assessment of ferning patterns has been proposed as a simple and economical test of dry eye [172178]. Normal individuals commonly display dense, uniform, closely branching ferns (Fig. 1) that change with altered tear functionality and chemical composition.

Fig. 1.

Fig. 1

Examples of tear ferning patterns for Rolando’s grade 1 (A), 2 (B), 3 (C), 4 (D). From Maragou et al., 1996. Clin Rheumatol. 15: 125–32 [177].

Collection of tears from the lower temporal tear meniscus with a capillary tube is most commonly used [124,126,178190], with a low coefficient of variance [191]. It should be borne in mind that the composition of tears collected from the meniscus may not be reflective of that over the entire ocular surface, and thus the ferning pattern obtained may differ with differing collection sites. A tear sample of 1–4 μL is expelled onto a clean glass slide, allowed to dry at stable room temperature (20–26 °C) and humidity (≤50%), observed using white light microscopy and photographed within 10–15 min. Factors shown to influence the ferning pattern are tear volume [173], room temperature and humidity during drying [190], covering the tear sample during drying [174], time between collection and assessment [174], and oil on glass slides [191].

Many authors [124,126,177,179,181183,186,188,190,192198] grade tear fern patterns using the 1 (complete, uninterrupted ferning pattern with no spaces between ferns) to 4 (complete absence of ferning) scale of Rolando (Fig. 1) [172]. The scale displays good intra- and inter-observer agreements, particularly when collapsing grades for normal (grade 1 or 2) and abnormal (3 and 4) patterns [185,199]. However, representation of only a portion of the entire ferning pattern [185], and usage of full grades [199] may have increased the repeatability of the scale.

Tear ferning patterns are independent of sex [175,181], hormonal fluctuations in women with a regular menstrual cycle [182,197], races on the Asian continent [181] and time of day during waking [180]. Degradation of tear fern patterns, indicating abnormal tear functionality, occur upon eye opening after wakening [179], with contact lens wear [186] and with increasing age, independent of dry eye status [124,181,198,200]. The effect of reflex tearing on ferning patterns remains equivocal, with Norn [191] demonstrating no effect on ferning patterns when reflex tearing was provoked by inhalation of ammonium chloride, but an improvement in ferning patterns after exposure to draught, which was hypothesized to result from reflex tearing [201].

Ferning in dry eye tears is less regular, with increasing space between ferns and shortening of ferns, culminating in an absence of ferning. Over 90% of keratoconjunctivitis sicca patients fell into ferning classes 3 and 4 [172,202] Sjögren syndrome patients exhibited 87.8% and 83.3% respectively with grades of 3–5 [176]. However, another study with Sjögren syndrome patients found only 59% and 48% showed class 3 and 4 ferning patterns [177]. 50% of post-menopausal women with dry eye symptoms fell into class 2 [124]. Using average grades, mean and median values range between 2.0 and 2.9 for participants with dry eye [126,176,182,195]. Using the 5-point Masmali scale, the authors reported a median value of 2.3, with 72.5% of samples graded between 2.0 and 3.0 [178]. Classifying ferning into normal and abnormal patterns or comparison of mean values between dry eye and non-dry eye participants can be analytically effective with this scale. Overall good sensitivity (80–90%) and specificity (75–89%) has been demonstrated for Sjögren syndrome [186], but more studies are needed to establish sensitivity and specificity and cut-off values in other forms of dry eye before this relatively simple approach is translated into private practice.

2.6. pH

Mean tear pH values range between 6.8 and 8.2, with individual values ranging between 5.2 and 8.6 [203216]. Yamada et al. [215] compared the pH values of collected tears to those measured on the eye and found an alkaline shift of 0.1 pH units in collected tears, which was thought to be caused by interaction of the tear fluid with air. As with all physical measures of the tear film, the collection site may influence the result obtained. Tears collected for pH assessment are typically those from the lower meniscus and may not be reflective of the pH over the ocular surface.

There seems to be no effect of sex [205,212], but age [210,212], time of day [204,210], eye closure [204,210], prolonged eye opening [208,210,213], blinking [208] and reflex tearing [208] have been shown to affect pH values. An alkaline shift of approximately 0.2 pH units can be observed between 10 and 50 year olds or between groups of under and over 40 year olds [210,212]. Slight acidification of the tear film by approximately 0.2 pH units can be observed after eye closure overnight or for 1 h [204,210]. Blinking causes a slight acidic shift, but overall the tear film shifts from acidic to alkaline throughout the day, with a difference of usually no more than 0.6 pH units [204,208,210]. An alkaline shift of 2.5 ± 0.6 pH units/min has also been observed with prolonged eye opening, with a maximum value of 9.3, however, the tear film reaches equilibrium after 30–60 s [208]. It is important to remember that pH is measured on a logarithmic scale such that for every one unit change in pH there is a 10-fold change in the concentration of hydrogen ions.

Investigations in dry eye using pH indicator paper showed no significant difference in pH mean values between a dry eye and control group (7.46 ± 0.24 and 7.45 ± 0.23 respectively), although 12% of participants with dry eye displayed a pH value over 7.7, compared with only 4% in the control group [214]. Norn [205], using a micro-electrode inserted into the inferior conjunctival fornix, found tears in participants with keratoconjunctivitis sicca to be significantly more alkaline compared to non-dry eye participants when both eyes were measured (0.2 pH unit shift), but not when only one was assessed (0.05 pH unit shift). Reporting a systemic error, with the second eye measured showing an alkaline trend, the authors judged their observation in dry eye to be unreliable [205]. Using a similar method, Khurana et al. reported a small alkaline shift of 0.1 pH units in participants with dry eye compared to non-dry eye ones [214]. Various methods have been used to measure tear film pH, each having their own limitations. To date, there is limited evidence for a change in pH values with DED.

2.7. Tear evaporation

Inter-blink tear evaporation may contribute to dry eye. Tear evaporation rate has been considered an indicator of tear lipid layer stability [217220], with higher tear evaporation rates associated with increased tear thinning, ocular dryness and discomfort [220222]. An early study on humans measured the evaporation rate using an instrument adapted from dermatology [223], but this technique involved touching the corneal surface, which is a stimulus for reflex tearing and likely resulted in an overestimation of evaporation rates. Non-invasive analysis in an environment with relative humidity between 30 and 40% revealed a tear film evaporation rate of 0.4–50 × 10−7 g/cm2/sec (Table 3) [2527,67,123,224242]. This very large range of evaporation rates may derive from use of different assays. Measurements undertaken with ventilated chambers may better mimic natural air flow on the ocular surface than closed preocular chambers [243], although fast measurement (<10 s) closed chamber devices may be appropriate. Currently there is no commercially available instrument dedicated to detection of tear film evaporation. Thus, researchers have either used in-house designed systems or modifications of closed chamber dermatological units [25,158,239].

Table 3.

Evaporation of tears from normal eyes.

Technique used Evaporation rate (mean ± SD) ( × 10−7 g/cm2/sec) Relative humidity (%) Reference
Open chamber with no temperature and humidity control 26.9 (±NR) NR Hamano et al., 1981 [226]
Closed ventilated chamber using resistance hygrometry 50 ± 16.6 50% Cedarstaff & Tomlinson 1983 [227]
Temperature and humidity sensors interconnected with constant humidity 4.1 ± 0.4 30% Rolando & Refojo 1983 [228]
Closed chamber modified dermatologic evaporimeter with two sensors placed at a known distance from ocular surface 12.5 ± 1.8 NR Trees & Tomlinson 1990 [229]
Closed chamber, temperature and humidity sensor 15.6 ± 3.8 40% Tsubota & Yamada 1992 [230]
Closed chamber humidity sensor 14.7 ± 6.7 30% Mathers 1993 [217]
Closed chamber humidity sensor 0.4 (±NR) 50% Craig & Tomlinson 1997 [67]
Ventilated chamber with constant airflow of known water content 4.1 ± 1.4 NR Goto et al. 2003 [232]
Closed chamber humidity sensor 10.8 ± 5.3 NR Thai et al. 2004 [236]
Closed chamber humidity sensor 15.1 ± 7.3 30% Guillon & Maissa 2008 [237]
Closed chamber humidity sensor 5.8 ± 2.8 NR Khanal et al., 2009 [123]
Quartz crystal humidity sensor 5.0 ± 2.8 18% Kojima et al., 2011 [234]
Closed chamber humidity sensor 0.04 ± 0.01 (μl/min/cm2) 30% Arciniega et al., 2011 [233]
Infrared thermography in a controlled adverse environmental chamber 26 ± 11.1 45% Petznick et al., 2013 [235]
Ventilated chamber with constant airflow of known water content 7.7 ± 6.0 30–50% (in room) Hirayama et al., 2013 [238]
Closed chamber humidity sensor 55.6 (95% CI: 40.8–75.8) (g/m2/h) 56% Rohit et al., 2014 [239]
Flow Evaporimeter 26 (±NR) 40% Peng et al., 2014 [25]
Ocular thermography 66.1 ± 21.1 (W/min) NR Yeo et al., 2016 [240]
Closed chamber humidity sensor 79 ± 33 (g/m2/h) 45% Alghamdi et al., 2016 [241]

NR = Not Reported.

The effect of age and sex on tear evaporation rates is equivocal. Some studies have shown that older age groups (>45 years) have 1.5 times higher evaporation rates than younger adults and women have evaporation rates approximately 1.4 times higher than men [244246], but other studies have failed to show any impact of age or sex [240,247]. Subjects who blink more than normal (likely being a consequence of evaporation) or have a larger exposed ocular surface have increased tear evaporation rates [246,248]. Low humidity increases evaporation [249251], with tear evaporation rates of healthy individuals increasing by >40% with a 10% reduction in relative humidity [249]. Increased tear evaporation rates in low humidity (5%) environments has been associated with reduced lipid layer thickness and tear film stability [251]. Meta analyses point to a 2–3 fold increase in tear evaporation rate as a diagnostic cut-off for DED, with a sensitivity of 45.5–61.2% and a specificity of 79.8–90.6% [252].

3. Biophysical studies of tears

3.1. Tear film evaporation and the lipid layer

Goggle-based chambers detect very low but dispersed tear evaporation rates of 0.14 ± 0.07 μL/min in the healthy eye, but 0.26 ± 0.16 μL/min for patients with meibomian gland dysfunction (MGD) [252]. Tear film thinning kinetics measured by interferometry suggests a higher tear evaporation rate of 1.137 μL/min in the healthy eye, which is a value similar to that of water [25,252]. The thickness of the lipid layer may not affect the evaporation rate unless it is very thin (less than 24 nm) or completely absent [2427,67]. Nonetheless, increased expression of meibum in normal eyes correlates with reduced evaporation in both healthy individuals and patients with DED [233,253]. An abnormal tear film lipid layer is associated with increased evaporation [2427,67]. Evaporation rates differ substantially between different studies, likely due to challenges in controlling environmental factors such as air flow and humidity [254]. Whether only evaporation from the ocular surface is being measured is also not always clear.

Such in vivo variability makes in vitro studies attractive [255]. Meibum films with physiological thickness do not retard evaporation in in vitro studies [255]. Evaporation found in studies using meibum films is in the range of 6–8% [225,256,257], or less in studies using artificial tear film lipid layers of up to 100 nm thick [253,254,257,258]. Only non-physiologically thick, 4 μm or ≥ 17 μm, meibum films can reduce the evaporation of water in laboratory experiments, but then by only 7 or 23% respectively [225,254]. Straight chain alcohols and acids and pure wax esters used alone can retard evaporation [259261]. Monolayers of tear film-like wax esters can retard evaporation when used at temperatures 3 °C lower than their bulk melting temperature [258,260,261]. However, meibum has a much more complex composition, structure and thermotropic behavior compared to the model films, and the melting temperature of meibum (~30 °C) [262264] is lower than the ocular surface temperature (35 °C). Recently in vitro studies with the Interfacial Dewetting and Drainage Optical Platform (i-DDrOP) on silicone hydrogel (SiHy) contact lenses have shown that the mucoaqueous subphase beneath thinner parts of meibomian films thins and dewets faster compared to that seen below thicker “islands” in the meibomian layer [265]. Therefore, it might be possible that meibum may locally be an effective barrier to evaporation. Suppression of tear evaporation by the lipid layer is thus poorly understood.

Three evaporation suppression hypotheses have been proposed: (1) the effect is based on a certain organization of the lipid layer, as proposed previously [266]; (2) the lipid layer is not solely the structure that is responsible for the retardation of evaporation, but specific protein or mucin (or a combination) interactions are also needed [267]. However, whole tears collected by capillary tube, apparently with all tear lipid and protein species and at least some mucins [17,19,21,268272], evaporate in vitro with water-like rates; (3) water evaporation from the ocular surface is not at all controlled by lipids and the lipid layer has other functions [257]. A thick duplex film enriched with unsaturated and branched chain lipids, as found in meibum, is a much less effective barrier to evaporation than tightly packed lipid monolayers [273].

If tears do evaporate at the same rate as water, eventually the concentration of secretory mucins and proteins in the lipid layer would increase, and approximate mucin near the surface of the cornea [274]. A dense polymer network of mucin has a high hydration “water holding” property and might suppress further tear evaporation [275]. It is also possible that mucins, together with proteins and polar lipids from tears, enhance spreading and structure of the tear film lipid layer. Thus all tear film key constituents may contribute to increased evaporative resistance.

3.2. Clinical observation and harvesting of tear film lipids

The tear film lipid layer is derived predominantly from meibum from the meibomian glands and spreads upward over the mucoaqueous layer [8] with blinking. Thinning of the lipid layer superiorly may create a rise in surface tension as a driving force for spreading [8]. Spreading may initially involve interaction between polar lipids and the mucoaqueous layer [276]. Polar lipids then carry nonpolar lipids [277]. In the normal eye, spreading of the lipid layer can be observed clinically by interference video microscopy as an upwardly moving front of horizontally disposed, colored fringes (Fig. 2). Fringes reflect topographic variations in thickness of the lipid layer and are thereby reflective of its intermolecular organization. The lipid layer is compressed during the down stroke of the blink and restored in the upstroke, with only a moderate disturbance of its organization between consecutive blinks [4,26]. Spreading is rapid at first (about 10 mm s−1), although lagging behind the upper lid [278]. Spreading slows and stabilizes after one second or more, with the interference pattern showing remarkable stability over the remainder of the blink [34,278]. Spreading is slower and less laterally arranged in dry eye patients [34,278], with reduced stability of the lipid layer fringes [279], suggesting that intermolecular stability is lost.

Fig. 2.

Fig. 2

Tear film lipid layer interferometry grading patterns. From Yokoi et al., Correlation of tear lipid layer interference patterns with the diagnosis and severity of dry eye. Am J Ophthalmol. 1996; 122: 818–24 [35]. A = Grade 1 (gray uniform), B = Grade 2 (gray non-uniform), C = Grade 3 (few colors non-uniform), D = Grade 4 (many colors non-uniform), E = Grade 5 (partly exposed corneal surface).

The dynamic nature of the lipid layer allows it to respond to the shear stress induced by the changing area of the air-tear interface during blinking. The lipid layer must be compressible during the down-sweep [280,281], but must spread rapidly after the up-sweep of the lid [26]. Polar lipids decrease the surface tension of the air-tear interface and increase the spreading rate of the lipid layer. Non-polar lipids increase the compressibility and stability of the lipid layer [281]. Imbalances in composition show as impaired viscoelastic properties.

Human meibum comprises ~90% of the tear film lipid layer, and is often investigated as the archetype for the tear film lipid layer. Commercially available clinical devices (such as Yoshitomi’s meibomian gland pressure forceps, or Arita’s meibum collector) [120,282] ensure collection of ≤0.5 mg meibum per donor, thus allowing for multiple repetitions of experiments. The different meibum collection procedures (by glass microcapillary tube, Dacron swab, cytology micro brush, or spatula) and inherent intra- and inter-individual variability may contribute to variations in the amount and composition of the meibum samples [283285].

The most representative tear film lipid layer mimic might be the lipid extract of whole human tears. However, to-date only one study has used it for in vitro analysis of surface tension [286], which is not surprising considering the difficulties in collecting a sufficient sample for reproducible experiments. Lipids extracted from deposits adsorbed on worn contact lenses have been used [287], although they contain higher amounts of phospholipids than meibum. Schirmer strips can be used to collect lipids [288]. The tear film contains ~8% of amphiphilic (polar) lipids [289291] that may be released from the lipid-binding protein lipocalin [292] from the mucoaqueous layer upon organic solvent extraction. These phospholipids are at very low levels in meibum [21,272], but as surfactants, if they are not bound to lipocalin, may have a significant role at the tear interface [255,285,293].

3.3. Surface tension of tears

The tear film lipid layer lowers the tear surface tension, thus sustaining the tear film’s high area-to-volume aspect ratio. Tear surface tension of healthy individuals is 43.6 ± 2.7 mN/m [294], rising to 53.0–55.5 mN/m after delipidation, which is remarkably close to the upper boundary of values observed in samples collected from patients with dry eye, of 49.6 ± 2.2 mN/m [286].

Supplementing delipidated tears with meibum lipids does not restore the surface tension [286,295]. Normal surface tension may thus rely on lipids and perhaps some proteins and glycoproteins from the mucoaqueous layer, such as lipocalin, lysozyme, and mucin, that are thought to intercalate with the lipid layer [272,276,296]. Increased surface tension in dry eye implicates abnormally enhanced protein content at the air/tear interface as a consequence of a compromised lipid layer. Proteins appear to take hours to penetrate a normal meibum lipid layer [297300].

Data on the thermotropic phase transitions (via spectroscopy, differential scanning calorimetry and X-ray diffraction studies) [262,264,301303] and interfacial and bulk rheology [29,301,302] of tear lipids in vitro and in vivo, suggest a thick viscoelastic duplex structure composed of a monomolecular layer of amphiphilic polar lipids at the mucoaqueous surface and a generally unstructured shear thinning lipophilic suspension at the lipid-air interface. The lipophilic suspension consists of lipid lamellar-crystallite particulates immersed in a continuous liquid phase with no long-range order. An alternative hypothesis rebrands the old folded trilayer model of McCulley and Shine and views the tear film lipid layer as a lamellar sandwich [304]. Lamellar models demand an ordered, almost rigid solid [302], which conflicts with the need for rapid reorganization during blinking.

Due to the abundance (≥90%) of non-polar lipids, both meibum and contact lens extracts do not spread as monolayers at the air/water surface, but instead form thick (10 nm to >100 nm) duplex films. They are thick enough to display bulk properties with two separate interfaces, but thin enough to neglect the effects of gravity) [257,285,288,301,302,305]. Uniform spreading of meibum is enhanced by polymers that contain polyanionic polysaccharide moieties, such as hyaluronic acid and secretory mucin glycoproteins [296,305], that form polymer interfacial gel-like networks due to hydrogen bonding of the polymer moieties with each other and with the phospholipid head groups. As a consequence, film viscosity is enhanced, and a more uniform 2-dimensional (2D) distribution of lipids and water is achieved [257,296,305].

Films of meibum and contact lens lipid extracts are non-collapsible, which helps them to withstand the dramatic area changes during the blink. In some studies, meibum layers show almost full surface pressure (π)/area reversibility (i.e. very low π/area hysteresis), which point to meibum’s capacity to rapidly reorganize during area cycling (as a model for the interblink period) [29,296,305307]. Interestingly, meibum films expand and thin at lower temperatures (23–25 °C), but shrink and thicken at physiological (≈35 °C) temperatures [29,263,301,305,308]. Such behavior agrees with the liquid suspension model of the non-polar lipids of the lipid layer. The non-polar lipid oily cap might act as a lipophilic solvent [29,301,309311] to accommodate polar lipids, which may form inverse micelles within it, during temperature-induced interface/bulk redistribution of the polar layer and during film compression. In other words, it may prevent expulsion of polar lipids into the aqueous phase such that polar lipids can promptly return to the mucoaqueous interface at film expansion.

Meibum interfacial properties and structure are largely insensitive to changes in osmolarity [312,313]. However, both meibum and contact lens lipid extracts are sensitive to the presence of (glyco)proteins (lactoferrin, lysozyme, mucin, lipocalin, serum albumins, lactoglobulin, lactoferrin, secretory IgA, keratin, lung surfactant proteins) [297300,307,314] and pharmaceutical agents (hyaluronic acid, benzalkonium chloride, SofZia, Polyquad, whole eye drops and lens care solutions) [305,315317] in the film subphase. However, even in sub-physiologically “diluted” (5–10 mN/m vs. ~30 mN/m) meibum films, the kinetics of the penetration of such substances is much slower (>1 h) than the physiological time scale [297300].

An alternative mechanism has been proposed, whereby the proteins or pharmaceuticals dissolved in the mucoaqueous subphase do not have to insert into the meibum or contact lens lipid extracts films. Instead, they need only to interact with polar lipid head groups aligned at the mucoaqueous interface. This is a rapid interaction that immediately alters the structure and dynamic interfacial properties of the layers [287,305,311]. These data, along with studies evaluating the impact of lipid (or lipophilic) inclusions directly into meibum layers, demonstrate that non-surface active or lipophilic substances are well tolerated by tear lipid films [318322], while polar surface active ingredients can disrupt the film’s structural integrity and surface properties [314317]. Interactions between the tear film lipid layer and pharmaceutical agents are of particular pharmacokinetic importance as they may impart long term effects at the ocular surface, in light of the much slower in vivo turnover rate of the tear film lipid layer compared to the mucoaqueous layer (0.93 ± 0.36%/min vs. 10.3 ± 3.7%/min respectively) [323].

Shear rheology experiments have demonstrated the viscoelastic nature of meibum films, which were able to stabilize tear film-mimicking films and to reduce the critical film thickness for dewetting [28]. The protective effect of meibum layers was stronger than that of Newtonian arachidonic acid and primarily viscous dipalmitoylphosphatidylcholine films, in line with the classic paradigm that viscoelastic surfactants are particularly effective at stabilizing thin films [324].

Meibum and contact lens lipid extract films also display visco-elastic behavior in dilatational rheology studies [2931], with predominantly elastic properties, thereby enabling resistance to deformation during tear film breakup. Whereas films of meibum collected from healthy donors are continuous, thick and predominantly elastic, films from patients with meibomian gland dysfunction are discontinuous, patchy and show viscoelasticity compromised at time scales (ie at low frequencies) encompassing the durations of blink and interblink intervals found in vivo [29]. The in vitro data correlate well with the impaired spreading and heterogeneous structure of the tear film lipid layer of patients with MGD seen in vivo [29]. Similar characteristic differences in the viscoelastic properties are reported between films of contact lens lipid extracts from Caucasians versus dry eye-susceptible Asians [30,325]. More attention should be paid to the capacity of the lipid layer to elastically stabilize the air/tear interface.

3.4. Tear film lipid layer bulk properties

Differential scanning calorimetry reveals that the phase transitions in meibum start at 10–15 °C and end at 35–36 °C, with a melting temperature (Tm) of ~30 °C [264,302]. Thus, it is reasonable to assume that in the meibomian gland orifices meibum is both disorganized and liquid to facilitate excretion. When exposed to the cooler (33–35 °C) ocular surface, meibum solidifies to a partially melted liquid crystalline state, thereby enhancing lipid layer in-plane elasticity, and via denser 2D molecular packing, ensuring a barrier to water evaporation.

Combining differential scanning calorimetry with small- and wide-angle X-ray diffraction points to meibum as a liquid suspension consisting of lipid lamellar-crystallite particulates immersed in a continuous liquid phase, with no long-range order at physiological temperature [262,264,301303]. Hot stage cross-polarized light microscopy confirms such a view and reveals that in meibum from patients with MGD there is an increased presence of non-lipid, non-melting, chloroform-insoluble inclusions of protein, including cytokeratin [262]. Spectroscopy studies also confirmed that MGD meibum contains more protein and relatively less methyl groups (CH3) and cis double bonds (cis=CH) compared to normal meibum [303]. The increased content of non-melting proteinaceous particles that are poorly miscible with lipid can account for the impaired spreading and discontinuous structure of meibum films in vitro and for their worsened viscoelasticity and evaporation suppression capability in vitro and in vivo [2931].

Bulk shear rheology evaluation suggests that meibum is a shear thinning liquid of extremely high viscosity [302]. At 35 °C, the shear viscosity of bovine and human meibum is approximately 105 greater than that of water and 3–4 orders of magnitude more viscous than mineral oil. However, such high bulk viscosity is inconsistent with the high mobility of fluorescent probes reported in meibum layers [296] and with the rapid reorganization of meibum films during area cycling [29,296,305307]. More research needs to be performed to resolve this issue.

Research on the bulk rheology of non-stimulated (or mildly stimulated) tears is relatively old, but in light of the new findings on the tear lipidome, there is a need for reappraisal. Human reflex tears collected after cold air stimulation are non-Newtonian and shear thinning [326]. Their viscosity at rest is ~9 cP and when shear is applied it rapidly drops to 1 cP (at shear rates ≥ 100 s−1) [327,328]. This indicates that the tear constituents, and particularly the compounds dissolved in the mucoaqueous layer, when at rest form a “transient” intermolecular network via weak non-covalent interactions (hydrogen-bonding, hydrophobic and/or electrostatic interactions etc.). Thus, they increase the mucoaqueous layer viscosity in the open eye, which raises its resistance to thinning and enhances tear film stability. At blink, the high shear applied by the eyelid breaks these weak interactions, the intermolecular network disintegrates and the aqueous can flow like water, thus preventing any damage that high viscosity may cause to the underlying corneal epithelium. The exact nature of the compounds involved in the transient intermolecular network remains unclear. Pure mono-component solutions of mucin or tear proteins in physiological concentrations are low viscosity and/or Newtonian fluids. Also, if the lipocalin-bound lipids are extracted, then tears become a low viscosity Newtonian fluid. This suggests that lipid/protein interactions, lipocalin and possibly lysozyme and lactoferrin play important roles in the shear thinning property of tears.

An interesting study found that when whole human tears are subjected to shear (at rates of 2–160 sec−1), dry eye tears need >10 times longer relaxation times compared to normal tears (2.8 ± 0.14 s vs. 0.26 ± 0.12 s) in order to equilibrate after the shear is ceased [328]. This suggests that a longer period should be needed for dry eye tear film, compared to a healthy one, to stabilize at the ocular surface after a blink. This demonstrates the potential importance of elasticity for tear film stability.

4. Biochemical properties of tears

4.1. The tear lipidome

In the normal eye, clear oil can be expressed from the meibomian orifices, which are located anteriorly to the mucocutaneous junction of the lid rim [329], by pressing on the glands through the lids. Expressibility is greatest nasally and least temporally [330]. An average amount of meibum stored in the meibomian glands is in the range of several hundred micrograms per eyelid [26]. The lid reservoir contains at least 30 times the amount of lipid present on the surface of the tear film (approximately 300 μg vs. 10 μg, respectively) [331,332]. Comparison of the lipids from meibum and whole tears [21,255,333] showed that the classes and ratios of lipids in whole tear samples were very similar to the classes and ratios of lipids and ratios in meibum. The only exception was phospholipids, the majority of which may be from another source.

Tears from the meniscus and meibum have been extensively studied to determine if changes in one or more components can be correlated with dry eye. Overcoming the technical difficulties involved with collection and processing small volumes without contamination has formed a major part of this journey, and therefore these studies have included developing techniques for collecting, identifying and quantifying both specific families of lipids and individual lipids. Techniques for analyzing meibum and tear lipid components have been extensively reviewed in previous publications [159,276,284,309,334]. Techniques employed to measure lipid composition each have limitations. Mass spectrometry (MS) is excellent for identifying lipid species, but until recently, has been weak in quantification. Nuclear magnetic resonance (NMR) is quantitative, however, single signals may derive from many different species of lipids. Only by a process of elimination can individual lipids be attributed to particular spectra.

Meibum is composed of approximately 95% non-polar lipids and 5% amphipathic lipids. Collection methods can alter the composition of the amphipathic lipids [282]. In humans, the non-polar component is composed of 30–50 mol% (ie the fraction of the wax esters compared to the total amount of lipid expressed as moles) of wax esters [284,309,334], 30–45 mol% of cholesterol esters [21,335,336] and a small percentage of triglycerides (~2%) [309]. The wax esters generally have an oleic acid component (C18:1) and the alcohol components vary from C18–C30. A feature of the cholesteryl esters is that they have very long acyl chains, predominantly C22:1–C34:1 [21,272]. In both cases, the acid groups can have odd numbers of carbons (such as C25) and this is unusual because fatty acids are normally synthesized using acetic acid (2C) as the building block. Other lipids found in meibum include free cholesterol, which makes up <0.5 mol% and phospholipids, which make up <0.01 mol% of meibum [21]. The predominant amphipathic lipid family found in meibum is the (O-acyl)-ω-hydroxy fatty acids (OAHFAs), which comprise about 4 mol% of total meibum [21,336,337]. The amount of OAHFAs and the low level of phospholipids imply that OAHFAs are the major surfactants in the tear film lipid layer, and although this has not been shown, it gathers support from in vitro experiments where OAHFAs have been shown to readily spread over large surface areas of an mucoaqueous subphase [320].

The range of mol% of total lipids for each major lipid class means that to correlate one of these main classes of non-polar lipids with dry eye requires a very large change relative to normal. In this regard, it should be noted that an earlier finding that normal, but not dry eye subjects, did not have cholesteryl esters [338] has not been confirmed in more recent studies [333,336].

Particular genetic profiles that correlate with particular ratios of lipids in tears or meibum have not been studied, but if there were such correlations, ethnicity does not appear to be a factor, because a population study of Asians found similar lipid ratios [336] to those found in other studies [21,272]. The ratios of lipids in meibum do not vary with age [339], but there are changes in certain polar lipids (not identified) in meibum with sex and age [340]. Whether or not the ratio changes for an individual as they develop dry eye has not been studied, although there is a relatively large variation in mol% lipid types among individuals [341]. Recently, using NMR spectroscopy, squalene was detected in meibum [318,342]. Squalene is of interest because it illustrates an example of a lipid that might play a protective role rather than a role in the formation of the tear film lipid layer. Squalene is anti-inflammatory and antibacterial and could have antioxidant properties by scavenging free radicals produced by UV rays [343].

A recent development in the field has been to compare meibum and the lipids of whole tears from the meniscus [21,341]. This centers on the concept that the lipid layer of the tear film might contain not only lipids derived from meibomian glands, but also lipids that have migrated from the mucoaqueous layer. To put this in perspective, the upper limits of total lipid concentration reported in whole tears collected from the meniscus are 380–580 pmol/μL (~0.5 mmol/L) [291,344], of which non-polar lipids comprise ~85mol%, OAHFAs ~4mol% and phospholipids ~10mol%. For individual subjects, main lipid classes in whole tears were in the same ratios as those found in meibum, except for phospholipids that derive from the meniscus, irrespective of the analytical technique used. The relative amount of phospholipid varies markedly between individuals from 1 to 30 mol% of total lipids; lysophosphatidylcholine is the most abundant species [341]. However, over a 3-day period, the ratios for an individual were the same [341].

Whether phospholipids contribute to lipid layer performance is an area of debate. Some argue for insufficient available phospholipid [309], with most of the intact phosphatidylcholine (but not lysophosphatidylcholine) bound to lipocalin. Others suggest that there is sufficient phospholipid in stimulated tears to form a monolayer [292]. Assuming 0.5 mmol/L (5 × 10−4 mol/L) total lipid and 10mol% phospholipid (5 × 10−5 mol/L) in whole tears and a total tear volume of 10 μL (10−5 L), then the total number of phospholipid molecules [345] available for lipid layer adsorption would be 3 × 1014 (5 × 10−5 × 10−5 × 6 × 1023). If one assumes a lipid layer surface area of 2 cm2 (or 2 × 1016 Å2), ~70 Å2 (2 × 1016 Å2/3 × 1014) should be available per surface phospholipid, which is identical to the area occupied by a single phosphatidylcholine molecule [345]. This calculation is based on the understanding that all phospholipids in 10 μL of tears have adsorbed to the tear film lipid layer and none is bound to lipocalin in the mucoaqueous layer. There is a strong case for a ‘perched tear film’, and if this were the case then the true ocular surface tear volume would be about 0.06 μL. Overall, it appears that phospholipids are insufficient to form a layer. The countering argument is that phospholipids collected from tears of the meniscus were already part of the lipid layer at higher concentration and so there are more than ample phospholipids to form a layer at the ocular surface.

Several studies have investigated whether the collection technique of whole tears from the meniscus affected the lipid profiles obtained [288,290,291]. Rohit and co-workers compared flushed & reflex tears with basal tears and found a relative mol% increase of phospholipids and free cholesterol [291], pointing to a preference for basal tears, and suggested care to avoid reflex tearing during collection. Moreover, reflex or flush tears contain very low levels of several lipids. Others found lipid profiles from capillary-collected basal tears and flushed tears to be almost identical [290], and that tears collected by Schirmer strips had similar profiles to these but with a greater amount of total lipid (~7.5 times). Two studies examined the lipid profile on different regions of the Schirmer strips that were used to collect tear samples. Using an IR method, Borchman and colleagues estimated that in the first 5 mm of the Schirmer strip, only 5% of the lipids were from the meibomian glands and also noticed that the lipid profile did not change over the first 15 mm [288]. Contrary to these findings, Lam and coworkers using a MS method, showed that the first 5 mm was composed of 15% phospholipids and approximately 80% was made up of meibomian lipids [290]. They also noticed that in the next 5 mm of the strip, the quantity of meibomian lipids halved, with few other lipids being detected and no signal for phospholipids. These results suggest that the use of Schirmer strips to collect tears may be not reliable.

In another study of mixed purpose, the main lipid classes found in meibum of wearers of contact lenses with and without dry eye were compared. The opportunity was also taken to compare the effect of different collection techniques (microcapillary tube, Dacron swab, cytology micro-brush, spatulas) and extraction techniques (immediate and following storage) [283] on the lipid profiles obtained. A cytology micro-brush was the most efficient at recovering phospholipids (50%), followed by the spatula (40%), microcapillary tube (38%), and Dacron swab (23%), and the authors concluded that the choice of collection device and extraction method plays a significant role in the detection of meibum lipids, particularly phospholipids [283]. This study also showed that contact lens wearers not suffering from dry eye were more likely to have inorganic phosphates present in their meibum samples compared with samples from contact lens wearers who had dry eye [283]. Another study that compared lipid composition of meibum collected using cotton buds, the meibomian gland evaluator and meibomian gland forceps demonstrated that the use of the evaluator resulted in smaller polar lipid quantities than the cotton bud approach [282]. There were no significant differences between techniques for nonpolar lipids [282].

Several studies have evaluated the impact of age on the composition of meibum and lipids from whole tears from the meniscus and compared these with subjects suffering from MGD [288,303,339,346350]. Interestingly, Lam et al. found OAHFAs to increase with age [333], which is unexpected because they also found that OAHFAs decreased with dry eye [336]. There was no difference in meibum phospholipid between males and females [351], but another study has shown a difference in polar (not identified) between the sexes [340]. Studies using NMR have shown that there was an increase in order of the molecules in MGD relative to age-matched controls and a corresponding decrease in lipid order with age [303]. An investigation of the meibum from Asians with different severities of dry eye showed that there was no overall difference in the ratios of the different classes of lipids between dry eye patients and normal subjects [336]. The proportion of total lipids that were triacyl glycerols was higher in the mildest category of dry eye compared with more severe categories, and the relative percentage of OAHFAs to total lipids progressively fell with increasing severity of dry eye. This was the only class of lipids to do so [336]. Some changes in the relative percentage of individual species of lipids to total lipid detected were found, but the trends were not consistent in direction across the dry eye categories.

Recently, several studies have profiled lipids in whole tears from the meniscus of patients with different severities of DED [21,289292,333,344,351,352]. Beyond the relative ease of tear collection versus meibum, marked levels of phospholipids are detectable in tears, ratios of lipids in whole tears from the meniscus reflect lipid-species ratios in meibum, and all meibum lipids have been identified in the tears. Table 4 summarizes recent studies (adapted from Millar and Schuett) [255]. Lam et al. observed a decrease in low molecular weight wax esters and a tendency of medium molecular weight wax esters to be more unsaturated although total tear lipids were unchanged in dry eye patients versus control subjects [290]. Another study by the same group used HPLC/MS/MS to determine correlations between structure-specific human amphipathic tear lipids and dry eye [333]. They found that several amphipathic lipids, cholesteryl sulfates, glucosylceramides, ganglioside mannoside 3, and lysophosphatidylcholines, were positively correlated with increased Schirmer I results, and that phosphatidic acids and phosphatidylglycerols showed a negative correlation. Small amounts of cholesteryl sulfate in meibum and tears from patients suffering with dry eye syndrome have been detected [290,333]. However, its role in meibum or the tear film needs further investigation. These components are commonly associated with epithelia, where they are believed to have a secondary messenger role [353]. It is possible that the cholesteryl sulfate detected in Lam et al.’s samples [290] may have originated from meibomian ductal cells.

Table 4.

Concentration of lipids in tears (adapted from Millar and Schuett) [255].

Lipid Saville et al., 2010 [289] Saville et al., 2011 [351] Rantamaki et al., 2011 [352] Dean and Glasgow 2012 [292] Brown et al., 2013 [21] Rohit et al., 2014 [291] Lam et al., 2014 [290] Lam et al., 2014 [333] Lam et al., 2014 [344]

Basal Reflex Flush

Collection method Capa Cap Cap Cap (reflex) Cap Cap Cap Cap Sch Sch Sch
Mol %
Cholesterol-Ester 39 54.8 35.7 33 44.8 43 33
Wax Ester 43 29.1 18.6 30 35.2 30 46
Cholesterol 8.2 17.2 ~20 5.9 8 8
OAHFA 4.4 ~2.5 <1 <1 2.5 2 1.1
Trigylceride 2.1 2.8 1 2
Phospholipids:
Phospholipid total 12
Phosphatidylcholine 1.6 2.5 2 3
Lysophosphatidyl-choline 8 0.5 0.8 1.7
Phosphatidyl-ethanolamine 0.8 0.7 1 1.2
Lysophosphatidyl-ethanolamine 1.3 1.6 2
Phosphatidylserine 0.5 0.1 0.5 0.3
Sphingomyelin 1.1 1.5 3
Ceramide

Concentration

pmol/μl ng/ng tear μM ng/ml pmol/μl μmol/ml

Cholesterol-Ester 178.1 27.9 13.5
Wax Ester 81.8 15.5 8.3
Cholesterol 16.6 9.8 5.9
OAHFA 4.3 0.4 0.1
Trigylceride 1.5 1.6 0.1 0.2
Phospholipids:
Phospholipid total 11 29.6
Phosphatidylcholine 6 20.9 194 1.6 1.1 0.5
Lysophosphatidyl-choline 3.5 9.5 3.2
Phosphatidyl-ethanolamine 5.2 0.1 0.1 0
Lysophosphatidyl-ethanolamine 1.5 1.2 0
Phosphatidylserine 0.1
Sphingomyelin 5 0.9 1.9 1.6 0.6
Ceramide 0.9
Total lipid (pmol/μL) ~380 ~100 ~40 0.58
a

Cap = collected by capillary tubes; Sch = collected by Schirmer strips.

In summary, tear lipid profiles from studies to date are variable, warranting investigation of the source(s) of variation (collection methods, analysis techniques or a combination of both) and close collaboration among different tear and meibum lipid laboratories worldwide. An initial step would be comparative analysis of a common shared sample towards comprehensive procedural standardization of all analytical and collection steps.

4.2. Mucins

Mucins are large high molecular weight glycoproteins that contain one or more protein domains, rich in serines and threonines, which are extensively glycosylated via O-glycan attachments. These important glycan chains constitute a significant part (50–80%) of the weight of the mucin [354,355]. Mucins, the building blocks of mucus, in general are present at apical epithelial surfaces of the respiratory, gastrointestinal, and reproductive tracts, as well as the ocular surface. Mucins provide a variety of protective functions, including but not limited to, lubrication, barrier formation and hydration. Fig. 3 demonstrates the role of mucins in the tear film structure.

Fig. 3.

Fig. 3

The tear film structure showing the mucins and galectin of the glycocalyx, soluble mucins and proteins in the mucoaqueous layer and the surface lipid layer. Adapted from Craig et al., 2013 [54].

At least 20 mucin genes (MUC) have been identified in humans [356], the products of which are classified into two types: secretory mucins and transmembrane mucins. The secretory mucins are further subdivided into large gel-forming mucins and small soluble (or non-gel forming) mucins [356]. The gel forming mucins are MUC2, MUC5AC, MUC5B, MUC6 and MUC19, the small soluble mucins are MUC7, MUC8 and MUC9, and 10 others have been characterized as transmembrane [356]. Ocular surface mucins are synthesized by corneal and conjunctival epithelia and lacrimal glands, where they contribute to the epithelial barrier and prevent binding of pathogens to the ocular surface, and are present in the tears where they maintain hydration [17,18,271,357359]. The role of mucins in ocular surface health and DED is a subject of great interest, that has been reviewed a number of times in recent years [15,359362].

Gel-forming mucins can be incredibly large, up to 40 MDa, thanks to their ability to multimerize through disulfide bonding of cysteine rich domains that flank the large central tandem repeat region [355]. The source of gel-forming mucins on the ocular surface are goblet cells of the conjunctiva, with MUC5AC being the most abundantly expressed [271,359] (though mRNA for others have also been found [363,364]). The abundance of MUC5AC present on the ocular surface is determined by the number of conjunctival goblet cells and the stimulation of its secretion by these cells. Gel-forming mucins are important for hydration of the ocular surface, and have also been suggested to play an important role in not only tear structure, but also in preventing debris and pathogens from binding to the ocular surface and damaging cells [18] by clearing them from the ocular surface through entrapment in the mucus layer and blinking [17,359].

Several studies have demonstrated decreased levels of MUC5AC in dry eye, yet it is important to note that altered levels of MUC5AC (and gel forming mucins in general), both decreased and increased, can negatively impact ocular surface health. As such, mucin production must be tightly regulated for ocular surface homeostasis. In patients with moderate DED a decrease in the total number of goblet cells, as well as empty goblet cells, has been observed by impression cytology [365]. MUC5AC has also been shown by impression cytology to be lower in dry eye patients with severe symptoms compared to mild and moderate symptoms [366], which was associated with increased inflammation, and (along with decreased levels of MUC16) in patients with an unstable tear film or aqueous deficiency compared to control eyes [367]. Consistent with these findings, very recently, goblet cell deficiency, and thus decreased MUC5AC expression, was observed in subjects (office workers using visual display terminals) with DED, particularly so in moderate to severe cases [368]. In patients with atopic kerato-conjunctivitis, the number of goblet cells and MUC5AC mRNA levels were significantly diminished [369]; a concomitant increase in MUC16 was also observed and the authors speculated that MUC16 compensates for lack of MUC5AC. In Sjögren syndrome patients, decreased MUC5AC mRNA and protein in tears and decreased MUC19 mRNA and protein has also been reported [270,364]. In summary, there is an abundance of data demonstrating DED is associated with decreased levels of MUC5AC.

MUC7 is currently the only small soluble mucin to be detected in ocular tissue, but it has not been detected in tears [271,363,370,371]. MUC7 is small (~180 kDa) [372] compared to gel-forming mucins, and exists as a monomer with a typical rigid structure ‘bottle-brush’ shape provided by water-binding O-glycans [361,373]. While MUC7 is expressed in the acinar cells of the lacrimal gland and in human conjunctiva, it is not known which cell type is responsible for synthesis and secretion in the conjunctiva, nor is it known what stimulates MUC7 secretion in ocular tissues [271,363,370,371]. As MUC7 has not been detected in the tears, its putative role in the tear film is not well understood. It is reasonable to speculate, based on MUC7 having the expected O-linked glycans of a mucin, that it would contribute to hydration of the ocular surface. However, no studies to date clearly demonstrate the function of MUC7 on the ocular surface. MUC7 expression, as assessed by mRNA levels in conjunctival impression cytology, has been reported to not be different between normal and patients with dry eye [374].

Transmembrane mucins expressed in the most superficial cell layer of the corneal and conjunctival stratified epithelia include MUC1, MUC4, and MUC16 [15,17,375]. MUC20 is one of the more highly expressed mucins in the human conjunctiva and is predominant along the plasma membranes of intermediate cell layers [376,377]. Transmembrane mucins have large extracellular domains, a hydrophobic membrane-spanning domain, and a short cytoplasmic domain. The extracellular domain comprises the tandem repeats rich in serine and threonine, which are subject to O-linked glycosylation [354], and can extend up to 200–500 nm above the ocular surface [362]. Soluble forms of MUC1, MUC4, and MUC16 have been detected in tears [271]. MUC16 has recently been localized to mucin granules of goblet cells as well, though its function remains to be determined [16]. In general, the trans-membrane mucins play important protective roles on the ocular surface. These include, consistent with those mentioned above, hydration of the ocular surface, and anti-adhesive properties between the cells of the conjunctiva and ocular surface [361], as well as contributing to tear film stabilization [15]. The cytoplasmic domain can associate with cytoskeletal components and also have numerous phosphorylation sites that suggest it may function in cell signaling [378,379]. A direct comparison between MUC 1 and MUC16 barrier function in the same epithelium suggested that transmembrane mucin barriers vary in the context of other trans-membrane mucins, and that MUC16 provides a major barrier when present [380]. Messenger RNA transcripts of the transmembrane mucins MUC13, MUC15, MUC17 have been identified in human conjunctiva [381], but neither their presence in tears nor their functions have yet been elucidated.

Transmembrane mucins also contribute to the highly organized protective epithelial barrier on the apical glycocalyx of the ocular surface epithelia cells though binding to galectin-3 (a multivalent β-galactoside-binding lectin) [382]. Indeed, galectin-3 co-localizes with MUC1 and MUC16 on the apical surface of epithelial cells, in a carbohydrate-dependent manner, and plays a key role in maintaining mucosal barrier function. The concentration of galectin-3 protein in tears (but not conjunctival epithelium expression) was found to be significantly higher in patients with dry eye and associated with epithelial dysfunction in dry eye, and the data demonstrate proteolytic cleavage of galectin-3 that may contribute to impaired ocular surface barrier function [383].

Changes in transmembrane mucin have been observed in patients with DED, but remain controversial. In post-menopausal women with non-Sjögren syndrome dry eye, increased MUC1 and MUC16 mRNA and protein levels have been reported relative to controls [384]. Similarly, increased MUC1 and MUC16 mRNA and soluble protein levels have been observed in Sjögren syndrome [385,386]. Conversely, decreased levels of MUC1 and MUC4 mRNA levels were reported in conjunctival epithelia of dry eye patients, and those authors suggested MUC1 mRNA levels could be used as a diagnostic biomarker for dry eye [374]. Such alterations could be the result of dry eye-related inflammatory mediators (eg IL-1, IL-6, IFN-γ, TNF-α) regulating the expression of transmembrane mucin, although, to date, studies examining this hypothesis have been in vitro only [387389]. In patients with complete androgen insensitivity syndrome who have DED [390], MUC1 mRNA and protein as well as the gel forming MUC5AC levels were decreased compared to those in control patients [376].

Altered glycosylation of ocular mucins has been observed in dry eye. Differential regulation of genes involved in glycosylation, as well as altered distribution of the responsible enzymes, has been observed in the conjunctiva of patients with DED [376,391]. Alteration of MUC16 sialylation, as shown by using the monoclonal antibody H185, has been detected in patients with dry eye symptoms [392,393]. Upregulated sialylation of MUC1 in mild-to-moderate dry eye and a decrease in severe dry eye has also been observed [394]. Indeed, altered sialylation of mucins in DED appears to depend on disease severity; whereas increased sialylation has been noted in the earlier stages of disease development, it is eventually lost with disease progression [395,396]. Increased sialylation on the cell surface would increase negative charge and enhance cell repulsion, as well as hydration through attraction of water. Additional work is required on the glycosylation of mucins and mucin-like molecules in DED.

A mucin-like glycoprotein called lubricin, or proteoglycan 4, is present on the ocular surface, where it provides lubrication and plays a protective role [397]. Lubricin was originally discovered in synovial fluid and on the surface of cartilage, and was found to also be present at the epithelial surface of human cornea and conjunctiva, though it is yet to be detected in tears [397]. Lack of lubricin on a mouse eye is associated with increased ocular surface damage, as demonstrated by an increase in corneal fluorescein staining [397], though to date no studies exist demonstrating alterations in lubricin expression with dry eye. In vitro, lubricin reduces friction at a human cornea-eyelid biointerface, as well as tissue-contact lens biomaterial interfaces [398400]. Recently, recombinant human lubricin was found to demonstrate improvement in both the signs and symptoms of DED compared to a control drop based on hyaluronate) [401]. Another heavily O-linked glycoprotein found in tears is Deleted in Malignant Brain Tumors-1 (also known as gp-340) and this has two isoforms [402]. The role of this protein in the tear film is unknown, although gp340 has been shown to be identical with the salivary agglutinin that strongly binds to bacteria and can activate the complement system [403,404] so may be involved in host defense.

In conclusion, mucins are critical to ocular surface health and deregulation of their synthesis is an important factor in ocular surface disease. Decreased MUC5AC expression and alteration in mucin glycosylation in dry eye is a consistent finding among most studies, while changes in transmembrane mucin expression is more variable (see Stephens & McNamara [360] for summary tables). Additional work is required to elucidate the role of individual mucins in disease, with the potential confounding factor of compensatory and/or complementary responses. Such work will likely benefit from the use of new, sophisticated genomic, glycomic and proteomic methods.

4.3. The tear proteome

The mucoaqueous layer contains salts and numerous proteins derived from the lacrimal gland and conjunctiva. A much expanded understanding of the tear proteome [405,406] has been achieved over the past decade and this section will concentrate on research conducted within this time frame. Techniques that have been used for proteomic analysis include unbiased tandem mass spectrometric (MS/MS) sequencing [19,268,269,407414], candidate immunocapture [368,415421] and other approaches (Table 5), from which a select group of putative disease biomarkers have emerged (Tables 6 and 7).

Table 5.

Mass spectrometric approaches.

Name Abbreviation Approach
Surface-Enhanced Laser Desorption/Ionization Time-Of-Flight Mass Spectrometry SELDI-TOF-MS Peptides are initially screened by differential binding chemistry (ie hydrophobic, anion, cation exchange), then are ionized and masses determined.
Matrix-Assisted Laser Desorption/Ionization Time-Of-Flight Mass Spectrometry MALDI-TOF MS Peptides are separated by liquid chromatography, or 2D or 1D polyacrylamide gel electrophoresis, then ionized and masses determined.
Isobaric Tag for Relative and Absolute Quantitation iTRAQ Stable isotope tagging to differentiate peptides from different sources, ie. normal vs. dry eye.
Tandem Mass Spectrometry MS/MS Ionized peptides are separated by mass, selected ions are fragmented, further separated and sequence derived.

Table 6.

Human ‘extracellular’ tear proteins reported to decrease or increase with ocular surface pathology. Right column: # articles detecting a significant change from normal by: unbiased mass spectrometric screening (‘US’) and/or a candidate immuno-detection or activity detection approach [CA]; NP, not performed. ‘Extracellular’, in whole or part by Gene Ontology.

Gene Symbol Protein Change from Normal & Pathology Known Function(s) # Articles US/[CA]
Angiogenesis
VEGF vascular endothelial growth factor A ↑ADDE [449], ↑MGD [417] angiogenesis, endothelial cell migration 0/[2]
ZG16B zymogen granule protein 16B ↓ADDE, ↓MGD↓, ADDE/MGD [414] angiogenesis, pancreatic adhesion, proliferation 1/[NP]
Calcium
AHSG alpha-2-HS-glycoprotein ↑SSDE [411] calcification inhibitor 1/[NP]
ANXA2 annexin A2 ↑SSDE [411] osteoclast formation and bone resorption 1/[NP]
ANXA5 annexin A5 ↓SSDE [411] promotes Ca2+ channel activity 1/[NP]
CALCA calcitonin related polypeptide alpha ↓ADDE [426] Ca2+ regulation 0/[1]
CALR calreticulin ↑SSDE [411] Ca2+ binding protein in ER and nucleus; SS assoc. 1/[NP]
NPY neuropeptide Y ↓SSDE [426] promotes Ca2+ signaling 0/[1]
Carbohydrate
ENO1 enolase 1, (alpha) ↑ADDE [407,414]
↑ADDE/MGD [414]
glycolytic enzyme 2/[1]
Carrier/Binding Protein/Steroid Associated
ALB albumin ↑ADDE [444,450], ↓BL [443], ↑CLDE [408], ↑FK [441], ↑MGD [451] carrier protein 4/[2]
AZGP1 alpha-2-glycoprotein 1 ↓ADDE [447,450], ↓MGD [409], ↑FK [441] zinc-binding, lipid degradation, cell adhesion 3/[1]
DMBT1 deleted in malignant brain tumors 1 ↓ADDE, ↓ADDE/MGD [414], ↑CLDE [408] scavenger receptor, binds surfactant protein D 2/[NP]
HP haptoglobin ↑ADDE [447], ↑CDK [445], ↑FK [441] hemoglobin binding, turnover to diminish iron loss 3/[NP]
LCN1 lipocalin-1 ↓↑ADDE [409,423,424,447,450], ↓FK [441], ↓MGD [451], ↓SSDE [411,428] hydrophobic protein binding, cysteine protease inhibitor 9/[3]
SCGB1D1 secretoglobin family 1D member 1a ↓↑ADDE [410,447], ↓CLDE [408], ↓MGD [451] in complex that binds steroids, including androgen 3/[1]
SCGB1D2 secretoglobin, family 1D, member 2 ↓MGD [451] in complex that binds steroids, including androgen 0/[1]
SCGB2A1 secretoglobin, family 2A, member 1b ↓↑ADDE [409,410,414,424,447], ↓ADDE/MGD [414] possibly binds steroids, including androgen 5/[NP]
SCGB2A2 secretoglobin, family 2A, member 2 ↓ADDE [447], ↑CLDE [408] cell migration? 2/[NP]
TCN1 transcobalamin I ↓ADDE [447] binds and helps move vitamin B12 into cells 1/[NP]
TF transferrin ↓ADDE [414,450], ↑ADDE/MGD [414], ↑SSDE [411] iron binding and transport to proliferating cells 2/[1]
Cell Adhesion/Motility/Structure
ACTB beta actin ↑MGD [409], ↑SSDE [411] cell structure, motility 2/[NP]
CFL1 cofilin 1 ↑SSDE [411] cytoskeleton organization 1/[NP]
LGALS3 lectin, galactoside-binding, soluble 3 ↑ADDE [446] cell adhesion, T cell regulation, antimicrobial 1/[NP]
LGALS7B galectin 7 ↓MGD [409] cell - cell and cell matrix communication 1/[NP]
MIF macrophage migration inhibitory factor ↑SSDE [411] chemo attractant and cell communication 1/[NP]
Cell Growth
EGF epidermal growth factor ↓↑ADDE [424], ↑MGD [417,431],↑ NK [431], ↓↑SSDE [430,431,451] mitogen: growth & differentiation 1/[6]
EGFR soluble epidermal growth factor receptor ↑ADDE [429] transmembrane receptor kinase 0/[1]
GCSF granulocyte colony stimulating factor ↑ADDE [449] colony-stimulating factor 0/[1]
GMCSF granulocyte monocyte colony stimulating factor ↑ADDE [449] colony-stimulating factor 0/[1]
LACRT lacritin ↓ADDE [447], ↓ADDE/MGD [414], ↓BL [443], ↓CDK [445], ↓CLDE [408], ↓FK [441], ↓SSDE [411,428,452] basal tearing, cell survival, wound healing 8/[1]
NGF nerve growth factor (beta polypeptide) ↑ADDE [426] neuronal differentiation and growth 0/[1]
Cytoprotective/Anti-Apoptotic
CASP14 caspase 14 ↓SSDE [411] apoptosis induction 1/[NP]
CLU clusterin ↑SSDE [411] inhibits apoptosis 1/[NP]
MUC5AC mucin 5AC ↓ADDE [368], ↓↑SSDE [270,411] mucus/gel-forming, cytoprotective, hydrophilic 1/[2]
PIP prolactin-inducible protein ↓ADDE [407,409,424,447],↓BL [443], ↑CLDE [408,453], ↓SSDE [411] inhibitor of T-cell apoptosis, aspartyl protease 8/[NP]
PROL1 proline rich, lacrimal 1 ↓ADDE [414,447], ↓ADDE/MGD [414], ↓MGD [414], ↓TAO [446] possible ocular protective function 3/[NP]
PRR4 proline rich 4 (lacrimal) ↓ADDE [410,414,442,447], ↓MGD [414] ↓CLDE [408], ↓SSDE [411,428], ↓TAO [446] possible ocular protective function 8/[NP]
Immune
C3 complement component 3 ↑ADDE [447], ↑SSDE [411] complement activation 2/[NP]
C4A complement component 4A ↑SSDE [411] cleaved to a trimer for complement activation 1/[NP]
CCL2 chemokine (C-C motif) ligand 2 ↑ADDE [429,449] monocyte, basophil specific chemotaxis 0/[2]
CCL3 chemokine (C-C motif) ligand 3 ↑ADDE [424], ↑SSDE [424] inflammatory response 0/[1]
CCL4 chemokine (C-C motif) ligand 4 ↑ADDE [424], ↑SSDE [424] inflammatory, chemokinetic 0/[1]
CCL5 chemokine (C-C motif) ligand 5 ↑ADDE [424], ↑SSDE [424] chemoattractant, histamine release 0/[1]
CCL11 chemokine (C-C motif) 11 ↑ADDE [449] Eosinophil chemotaxis 0/[1]
CCL15 chemokine (C-C motif) ligand 15 ↑ADDE [429] neutrophil, monocyte, and lymphocyte chemotaxis 0/[1]
CSF3 colony stimulating factor 3 (granulocyte) ↑ADDE [429] production, different, function of granulocytes, macrophages 0/[1]
CX3CL1 chemokine (C-X3-C motif) ligand 1 ↑ADDE [429], ↑MGD [417] neutrophil chemotaxis 0/[2]
CXCL5 chemokine (C-X-C motif) ligand 5 ↑ADDE [429] inflammatory cytokine, neutrophil activation 0/[1]
CXCL8 chemokine (C-X-C motif) ligand 8 ↑ADDE [427,449] inflammatory mediator 0/[2]
CXCL9 chemokine (C-X-C motif) ligand 9 ↑SSDE [436] T cell chemotaxis 0/[1]
CXCL10 chemokine (C-X-C motif) ligand 10 ↑SSDE [436], ↑MGD [417] T cell, monocyte chemotaxis 0/[2]
CXCL11 chemokine (C-X-C motif) ligand 11 ↑ADDE [436], ↑SSDE [436] T cell chemotaxis 0/[1]
ELANE neutrophil elastase ↑SSDE [411] phagocytosis and leukocyte migration 1/[NP]
IL1B interleukin 1 beta ↑ADDE [423,427,429,449],↑CCH [415] pro inflammatory response 0/[5]
IL1RN interleukin 1 receptor antagonist ↑ADDE [435,449], ↑MGD [417] cytokine activity 0/[3]
IL1R1 interleukin 1 receptor, type I (soluble) ↑ADDE [429] immune mediator 0/[1]
IL2 interleukin 2 ↑ADDE [427] T, B cell proliferation 0/[1]
IL4 interleukin 4 ↓ADDE [427,429] immune response, growth factor 0/[2]
IL5 interleukin 5 ↑ADDE [423,427] B cell, eosinophil growth & differentiation 0/[2]
IL6 interleukin 6 ↑ADDE [423,425,427,429,432], ↑CCH [415], ↑MGD [425], ↑SSDE [411,425] B cell, nerve cell differentiation 1/[7]
IL6R interleukin 6 receptor alpha ↑ADDE [429] regulation of immune response, hematopoiesis., acute-phase reaction 0/[1]
IL6ST interleukin 6 signal transducer ↑ADDE [429] shared interleukin transducing receptor subunit 0/[1]
IL-9 interleukin 9 ↑ADDE [449] Cell proliferation, anti-apoptosis 0/[1]
IL10 interleukin 10 ↑ADDE [427] inflammation, immunoregulation 0/[1]
IL12 interleukin 12 ↑ADDE [449], ↓ADDE [429] activated T, NK cell mitogen 0/[2]
IL15 interleukin 15 ↑ADDE [449] cell proliferation of natural killer cells 0/[1]
IL16 interleukin 16 (lymphocyte chemoattractant factor) ↑ADDE [429] CD4+ lymphocyte, monocyte, eosinophil migration 0/[1]
IL17A interleukin 17A ↓ADDE [429] proinflammatory response 0/[1]
IFNG interferon, gamma ↓↑ADDE [423,427,429] immune-regulatory, antiviral 0/[3]
JCHAIN joining chain of multimeric IgA and IgM ↑CDK [445] linkage of IgA or IgM monomers 1/[NP]
IG Immunoglobulin G ↑RA + SS [454] Immunoglobulin 0/[1]
ORM1 orosomucoid 1 ↑ADDE [407,414], ↑ADDE/MGD [414], ↑SSDE [411] modulator of acute-phase immune activity 3/[NP]
ORM2 orosomucoid 2 ↑SSDE [411] modulator of acute-phase immune activity 1/[NP]
PEBP1 Phosphatidyl-ethanolamine binding protein 1 ↑ADDE [414], ↑ADDE/MGD [414] modulates NF-kB in Sjögren syndrome 1/[NP]
SERPING1 serpin peptidase inhibitor clade A member 1 ↑SSDE [411] complement activation regulator 1/[NP]
TGFB activator transforming growth factor beta 1 & 2 activation ↑SSDE [420] suppress IL-2 T cell growth 0/[1]
TNF tumor necrosis factor ↑ADDE [423,427,432],↑SSDE [432] autoimmune, pyrogen, mitogen, differentiation 0/[3]
TNFRSF1B soluble TNF receptor superfamily member 1B ↑ADDE [429] part of TNF receptor heterocomplex 0/[1]
Lipid/Cholesterol
APOA2 apolipoprotein A-II ↑FK [441], ↑SSDE [411] in high density lipoprotein particles 2/[NP]
Other/Unknown
B2M beta-2-microglobulin ↑ADDE [410], ↓CLDE [408] loss associated with hypercatabolic hypoproteinemia 2/[NP]
POTEI POTE ankyrin domain family member I ↑TAO [446] ? 1/[NP]
SMR3B submaxillary gland androgen regulator protein 3B ↓ADDE [442] ? 1/[NP]
SOD1 superoxide dismutase 1 ↑SSDE [411] Antioxidant 1/[NP]
Phosphatase/Kinase/GTPase/Other Enzyme Phosphatase
F2 coagulation factor II (thrombin) ↑SSDE [428] fibrinogen to fifbrin conversion 1/[NP]
MMP9 matrix metallopeptidase 9 ↑ADDE [415],↑BL [415],↑CCH [415], ↑MGD [433], ↑PLDE [421],↑SSDE [433] matrix collagen IV and V degradation 0/[3]
S100A8 S100 calcium binding protein A8 (calgranulin A) ↑ADDE [407,409,410,414,424,442,446], ↑SSDE [411], ↑MGD [414,448], ↓TAO [446] cytokine & inhibitor of casein kinase/Inflammation 9/[NP]
S100A9 S100 calcium binding protein A9 (calgranulin B) ↑ADDE [407,409,414,424], ↑ADDE/MGD [414] ↑SSDE [423], ↑MGD [414,448] possible inhibitor of casein kinase 6/[NP]
Protease inhibitor/Antimicrobial
AZU1 azurocidin 1 (cationic antimicrobial protein 37) ↑SSDE [411] antibacterial and monocyte chemoattractant 1/[NP]
CST1 cystatin SN ↑ADDE [423], ↓MGD [409] cysteine protease inhibitor 1/[1]
CST4 cystatin S ↓ADDE [447], ↓BL [443], ↓FK [441], ↓MGD [409], ↓SSDE [428] cysteine protease inhibitor/antimicrobial 5/[NP]
CTSB cathepsin B ↓SSDE [428] lysosomal cysteine protease 1/[NP]
CTSS cathepsin S ↑SSDE [455] lysosomal cysteine protease 0/[1]
DCD dermcidin ↑SSDE [411] C-terminal antibacterial, N-terminal prosurvival 1/[NP]
IGHA1 immunoglobulin heavy constant alpha 1 ↓SSDE [411] microbial and foreign antigen defense 1/[NP]
IGKV3–20 immunoglobulin kappa variable 3–20 ↓BL [443] microbial and foreign antigen defense 1/[NP]
IGLC2 immunoglobulin lambda constant 2 ↓ADDE [447] microbial and foreign antigen defense 1/[NP]
LTF lactoferrin ↓ADDE [409,424,444,450], ↑FK [441], ↓MGD [451], ↓SSDE [411,430,433] iron metabolism, antibacterial 6/[4]
LYZ lysozyme ↓ADDE [407,447,450], ↓BL [443], ↓SSDE [411] hydrolase, antibacterial 5/[NP]
PIGR polymeric immunoglobulin receptor ↓ADDE [447], ↓CDK [445], ↓SSDE [411] antibacterial, polymeric Ig transcellular transport 3/[NP]
PLG plasminogen ↑SSDE [434] activates urokinase-type plasminogen activator, collagenase 1/[NP]
S100A7 s100 calcium binding protein A7 (psoriasin) ↑SSDE [411] chemotactic, antibacterial 1/[NP]
SERPINA1 serpin peptidase inhibitor clade A member 1 ↑ADDE [423,442], ↓BL [443], ↑FK [441] serine protease inhibitor, anti-inflammatory 3/[1]

ADDE = aqueous deficient dry eye; BL = blepharitis; CCH = conjunctivochalasis; CDK = climatic droplet keratopathy; CLDE = contact lens-related dry eye; FK = Fusarium keratitis; MGD = meibomian gland dysfunction; NK = neurotrophic keratopathy; PLDE = post-LASIK dry eye; RA + SS = rheumatoid arthritis with Sjögren syndrome; SSDE = Sjögren syndrome-associated dry eye; TAO = thyroid-associated orbit pathology.

a

Also known as lipophilin A.

b

Also known as mammaglobin B, mammaglobin 2 & lipophilin C.

Table 7.

Human ‘intracellular’ tear proteins reported to decrease or increase with ocular surface pathology. Right column: # articles detecting a significant change from normal by unbiased mass spectrometric screening (‘US’) and/or a candidate immuno-detection or activity detection approach (‘CA’); NP, not performed. Designated as ‘intracellular’ or associated with the ‘plasma membrane’ by Gene Ontology.

Gene Symbol Protein Change from Normal & Pathology Known Function(s) # Articles with Change from Normal: US/[CA]
Regulation of transcription/Nucleic acid metabolism
BAZ2B bromodomain adjacent to zinc finger domain 2B ↑ADDE [444] likely transcription regulator activity 1/[NP]
MDN1 midasin AAA ATPase1 ↑TAO [446] ribosome maturation 1/[NP]
RREB1 ras responsive element binding protein 1a ↑ADDE [444] zinc finger transcription factor 1/[NP]
SMARCA4 Transcription activator BRG1 ↑ADDE [446], ↓TAO [446] transcription regulation 1/[NP]
Cell communication/Signal transduction
ANXA11 annexin A11 ↑ADDE [409], ↑SSDE [411] calcium-dependent phospholipid binding protein 2/[NP]
PLAA phospholipase A2-activating protein ↑ADDE [409] secosteroid 1α,25-dihydroxy vitamin D3 signaling 1/[NP]
S100A6 S100 calcium binding protein A6b ↑ADDE [409], ↑SSDE [411] calcium binding/epithelial integrity, growth 2/[1]
S100A11 S100 calcium binding protein A11c ↑ADDE [407], ↑SSDE [411] calcium ion binding 2/[NP]
S100A12 S100 calcium-binding protein A12d ↑SSDE [411] proinflammatory, calcium binding 1/[NP]
TGFA transforming growth factor alpha ↑ADDE [429] targets EGFR to stimulate cell differentiation and proliferation 0/[1]
Cell growth and/or maintenance
S100A4 S100 calcium binding protein A4 ↑ADDE [407,409] calcium binding possibly in cell motility, invasion 2/[1]
Cytoprotective/Anti-Apoptotic
ALDH3A1 aldehyde hydrogenase 3A1 ↑ADDE/MGD [414] promotes survival of corneal epithelial cells 1/[NP]
HSPB1 heat shock protein family B (small) member 1 ↑ADDE [446] stress resistance, organization of actin 1/[NP]
Metabolism/Energy pathways
ANXA1 annexin A1 ↑ADDE [409,424,446],↑SSDE [411] inhibits phospholipase a2, anti-inflammatory 4/[NP]
GSTP1 glutathione S-transferase pi 1 ↑ADDE [409], ↑CCH [472] links reduced glutathione to electrophilic & hydrophobic molecules 2/[NP]
PRDX5 peroxiredoxin 5 ↑ADDE [409], ↑CCH [472] antioxidant, reduces hydrogen peroxide and alkyl hydroperoxide 2/[NP]
PLA2G2A phospholipase A2 group IIA ↑ADDE [470] hydrolysis of phosphoglycerides 0/[1]
Transport
AQP5 aquaporin 5 ↑ADDE [430], ↑SSDE [430] tear, saliva, pulmonary secretion; water channel protein 0/[1]
Immune response
IL1A interleukin 1 alpha ↑ADDE [425,433], ↑MGD [433], ↑SSDE [433,464] hematopoiesis, immune response, inflammation 0/[3]
IL33 interleukin 33 ↑ADDE [429] Th2 cell mat; mast/natural killer cell, basophil, eosinophil activator 0/[1]

ADDE = aqueous deficient dry eye; CCH = conjunctivochalasis; MGD = meibomian gland dysfunction; SSDE = Sjögren syndrome-associated dry eye; TAO = thyroid-associated orbitopathy.

a

Also known as zinc finger motif enhancer-binding protein 1.

b

Also known as calcyclin.

c

Also known as calgizzarin.

d

Also known as calgranulin C.

MS/MS runs are fed by trypsinized tear protein fragments that are first subjected to reverse-phase C-18 (without or with cation exchange [269,407]) separation. In some cases, tear proteins were first partitioned by 1- or 2-D sodium dodecyl sulfate polyacrylamide gel electrophoresis [269,408,409,414], which offers the advantage of separating monomers from polymers, or were collected into normal or dry eye pools for differential labeling by mass [407] or by fluorescence [408]. In other cases, tear proteins were first adsorbed onto protein chips, disks or tips bearing C-8, C-18 or cation ion exchange chemistry [410,422], or abundant albumin was first immunodepleted from tears [409], the latter an acknowledgement of the limited dynamic range of MS/MS, by which changes in less abundant tear proteins are overlooked.

Although well-suited for discovery, MS/MS coupled with chromatography is currently not practical as a quantitative, high throughput tear diagnostic, despite some success in a few academic studies. For this reason, candidate immunological approaches have been explored (Table 6) [270,368,415,417,420,421,423436]. These include antibody array (16 inflammatory cytokines [437]), sandwich ELISA (for MMP-9 [415,418,421], MUC5AC [368], IL-1β and IL-6 [415]), direct ELISA (lacritin [416,438]), and bead adsorption [439] (13 inflammatory cytokines, EGF, vascular endothelial growth factor (VEGF) [417,419], TGFβ1 and 2 [420]). An expanded selection of quantitative immunoassays built on validated biomarkers (Tables 6 and 7) should be the way of the future. Investigators should be careful to control for non-specific tear matrix effects [437,440], and independently validate commercial immunoassays where possible (one commercial lacritin ELISA displayed no reactivity for purified recombinant lacritin [405]).

Currently, almost 1800 proteins are known to constitute the human tear proteome. Before the original TFOS DEWS report in 2007, 491 proteins had been identified [268]. In 2012, Zhou et al. [19] reported 1543 proteins, of which 1304 were new. Over half are indicated to be ‘intracellular’ or associated with the ‘plasma membrane’ by Gene Ontology, but those that selectively decrease or increase in ocular surface diseases and thus are potential biomarkers are to date largely ‘extracellular’ according to Gene Ontology [270,368,407,408,410,411,425,428,430,431,434,441448].

Other technical considerations include the method of tear collection, whether topical anesthesia is employed to reduce reflex tearing [442], sampling from open versus closed eyes [268,437] and the optimal collection point on the eye. Capture of tears by microcapillary tubes is most common in proteomic studies [268,269,408,409,418421,432,437,450], although others collect tears onto Schirmer strips [269,407,410412,442], using sponges [409,415,429] or via micropipette tip [368,417,450]. Posa et al. [456] and Grus et al. [442] note that microcapillary tubes and Schirmer strips are equally effective, whereas Green-Church et al. [269] suggest that the tear film collection method impacts proteins present in the sample. However, Schirmer strips are preferred by patients for comfort and perceived safety. Schirmer strips also appear to retain proteins <40 kDa [442]. Most studies fail to mention the use of anesthesia, whether tears are from open or closed eyes, and the site of collection by tubes and sponges (sampling is slightly more common from the lateral [417419,450] versus medial [420] canthus). Noting whether anesthesia has been used is important, with relative levels of 15 tear proteins, possibly more, differing between basal and reflex tears [457]. Previous studies have shown that certain tear proteins such as lysozyme, lactoferrin, and lipocalin-1 are regulated, in that their concentration remains approximately the same during changes to the amount of tears being produced (for example during reflex tearing) [457]. Other proteins, such as sIgA, are constitutive or have a constant level of production, such that their concentration decreases during increases in tear fluid production [458]. Some proteins appear in the tear film via leakage from conjunctival blood vessels [459,460]. The concentration of sIgA could be used to normalize protein levels that might be affected by collection of reflex vs. non-reflex tears [461], and this is an area worthy of further study, although sIgA is deregulated in some pathological conditions. Indeed, it is currently not known whether most of the proteins within the proteome are constitutive, regulated or derived from serum (or elsewhere). A correct normalization process usually involves a “normalizer protein” or “housekeeping protein” [462]. A strict definition of a housekeeping protein is a constitutive protein that should be expressed under normal and pathological conditions at relatively constant levels. Such a protein has yet to be described in tears. A publication of cerebral spinal fluid proteomics indicates that the preferable way to quantify proteins is to first match the samples on total protein amount and then normalize the data based on the median intensities [463]. This technique has yet to be used in tear film biochemistry studies.

Human tear biomarkers might predict, diagnose and even, in some cases, be therapeutic for ocular surface disease. Nine early candidate biomarkers were identified primarily by immunoassay (see Tables 5 and 6 for expanded details):

  1. epidermal growth factor (EGF): reduced in Sjögren syndrome and aqueous deficient dry eye (ADDE) [430,464], augmented in evaporative dry eye due to MGD [417].

  2. interleukin 1 (IL-1α): increased in Sjögren syndrome dry eye and MGD [433].

  3. interleukin 6 (IL-6): increased in Sjögren syndrome dry eye [432,465].

  4. lactoferrin (LTF): reduced in Sjögren syndrome and non- Sjögren syndrome dry eye [466].

  5. lipocalin 1 (LCN1): a putative auto-antigen for Sjögren syndrome [467] that is decreased in MGD [468].

  6. matrix metalloproteinase 9 (MMP-9): increased in Sjögren syndrome dry eye, but unchanged from normal in MGD [433].

  7. MUC5AC: reduced in Sjögren syndrome dry eye [270,469].

  8. plasmin activity from plasminogen (PLG): increased in Sjögren syndrome dry eye [434].

  9. group IIA phospholipase A2 (PLA2G2A): increased in dry eye [470,471].

Seven more have been identified by unbiased MS/MS analysis of 2-D gel protein bands that differed in quantity between normal and blepharitis tears [443]. Selectively down-regulated biomarkers were albumin (ALB), immunoglobulin kappa variable 3–20 (IGKV3–20), serpin peptidase inhibitor clade A member 1 (SERPINA1), prolactin-induced protein (PIP), cystatin S (CST4), lacritin (LACRT) and lysozyme [443].

Topping the ‘extracellular’ list of deficient or augmented biomarkers in unbiased MS/MS screens (Table 6) in decreasing order of replications (respectively ≥6, 5, 4, 3, 2) listed alphabetically are (≥6) lacritin, lactoferrin, lipocalin 1, prolactin-inducible protein, proline rich 4 (PRR4) and S100 calcium binding protein A8 (S100A8); next (5) are cystatin S (CST4), lysozyme (LYZ) and secretoglobin family member 2A member 1; followed by (4) albumin (ALB) and S100 calcium binding protein A9 (S100A9); and (3) alpha-2-glycoprotein 1 (AZGP1), apolipoprotein A-II (APOA2), beta 2 microglobulin (B2M), haptoglobin (HP), orosomucoid 1 (ORM1), polymeric immunoglobulin receptor (PIGR), proline rich lacrimal 1 (PROL1), secretoglobin family member 1D member 1, serpin peptidase inhibitor clade A member 1; and finally (2) actin beta (ACTB), complement component 3 (C3), deleted in malignant brain tumors 1 (DMBT1), enolase 1 (alpha; EN01), secretoglobin family member 2A member 2, transferrin (TF). Thirty-four other ‘extra-cellular’ proteins are deficient or augmented in single unbiased MS/MS screens - many from Sjögren syndrome.

Among intracellular proteins (Table 7) [407,409,411,424,425, 429,430,433,444,446,464,470,472] in decreasing order of replications (respectively 4, 3, 2) are annexin A1 (ANXA1), followed by annexin A11 (ANXA11), glutathione S-transferase pi 1 (GSTP1), peroxiredoxin (PRDX5), S100 calcium binding protein A4 (S100A4), S100 calcium binding protein A6 (S100A6) and S100 calcium binding protein A11 (S100A11). Seven others are deregulated in single unbiased MS/MS screens to date. The surprisingly large representation of intracellular proteins in tears is presumed to be due to normal shedding of ocular surface epithelial cells [268], but why so few are more abundant in diseased tears is unknown. The classification of proteins by gene ontology, whilst useful, is a fairly blunt technique as, for example, secretory phospholipase A2 group IIA is classified as an intracellular protein when it is known to be secreted onto mucosal surfaces [470,473].

Not detected by MS/MS, but apparent by candidate immuno-screening or activity assays (Table 6) [415,417,420,421,423,424,426, 427,429,432,433,449,450] were several growth factors and cytokines. VEGF levels were increased in MGD [417]. Elevated in ADDE were soluble epidermal growth factor receptor, colony stimulating factor 3 (CSF3), transforming growth factor activity, and soluble tumor necrosis factor superfamily member 1B (TNFRSF1B) [429]. Levels of twelve chemokines were augmented in ADDE or MGD [424,427,429]. Fifteen interleukins or interleukin-associated molecules were for the most part increased in ADDE but two decreased and one increased in MGD [415,417,423,425,427,429]. There was disagreement over whether interferon gamma (IFNG) is increased or decreased in ADDE (although it is clearly increased in animal dry eye models) [423,427,429,474,475]. MMP-9 was increased in ADDE, post-LASIK dry eye, blepharitis, and conjunctivochalasis [415,421,433]. Five intracellular or plasma membrane tear proteins were detected only by candidate immuno-screening: transforming growth factor alpha (TGFA), phospholipase A2 group IIA (PLA2G2A), aquaporin 5 (AQP5), interleukin 1 alpha (IL1A) and interleukin 33 (IL33) - all increased in ADDE. AQP5 and IL1A were also increased in Sjögren syndrome, and the latter also in MGD [410,417,423,425,427,429,430,432,433,464,476,477]. It is noteworthy that of extracellular biomarkers most are immune-related (34 proteins), or fall under the category of protease/inhibitor/antimicrobial (14 proteins), or carrier/binding/protein steroid associated (11 proteins), or cytoprotective/anti-apoptotic (6 proteins). Intracellular ones are most commonly associated with cell communication/signal transductions (6 proteins), metabolism/energy pathways (4 proteins) or regulation of transcription/nucleic acid metabolism (4 proteins).

To date, half of the candidate biomarkers (Tables 5 and 6) lack immunological validation. Once confirmed, dry eye versus normal protein concentrations are needed, for which there should be good agreement among different laboratories. Initial values have been documented for several members of the interleukin family (IL1A, IL1B, IL-2, IL-4, IL-5, IL-6, IL-10), CXCL8 [415,424,425,427], TNF-α and IFN-γ [427,432], MMP-9 [415] and MMP-9 activity [421,433], LTF and EGF [425,430,433,450], ENO1, S100A4 [407], LYZ, TF, LCN1, ALB, AZGP1 [450], CALCA, NPY, NGF, CCL3, CCL4, and CCL5 [424,426]. Values for IL1B, IL-6, MMP-9 and TNF-α vary greatly. Most have not been subjected to replication by others. If the assumption that biomarkers repeatedly deficient or augmented in unbiased MS/MS screens may indeed be the most promising, then future priority could be given to the extracellular biomarkers that have been commonly found by MS/MS screening and, with the exception of LTF and LYZ, have not been subjected to this kind of analysis.

Poor efficacy of anti-inflammatory agents for dry eye highlights the need for new therapeutic strategies [478]. Some biomarkers -all extracellular ones deficient in dry eye - have shown benefits for ocular surface homeostasis and/or tearing in preclinical and, in some cases, clinical situations. Albumin [479], lactoferrin [480] and lacritin [481] each have prosurvival and wound healing activities [479,482]. When applied topically, albumin was reported to provide symptom relief in most dry eye patients with graft versus host disease [483]. Lactoferrin can promote healing of damaged human corneal epithelial cells in culture [484]. In a mouse corneal alkali injury model, topical lactoferrin protected corneal epithelial cells from hydrogen peroxide oxidative damage [485], and when fed to one year old mice, reportedly reduced inflammation of both lacrimal glands and the conjunctiva, and reduced oxidative stress [486]. A lactoferrin supplement was recently reported to increase tearing in a randomized double blind placebo controlled dry eye study in humans [487], and to restore tearing and TBUT in a randomized controlled study of cataract surgery-induced dry eye [488]. Topical lacritin reduced corneal staining in the Aire−/− mouse model of dry eye [489] and promoted the health of cultured human corneal epithelial cells exposed to interferon gamma and tumor necrosis factor [481] or to benzalkonium chloride [490]. Normal human tears immunodepleted of lacritin lacked cell survival activity. Spiking lacritin, but not the truncation mutant C-25, into ADDE tears restored cell survival activity [481]. Topical lacritin, but not C-25, promoted basal tearing when applied topically to eyes of normal rabbits [491]. Topical lacritin (discovered out of an unbiased screen for factors that promote basal tearing) [438] also restored tearing in Aire−/ − dry eye mice [489]. Tear film stability is enhanced by interaction of lysozyme with the tear protein vitelline membrane outer layer protein 1 homolog (VMO1) [492].

A successful biomarker(s) should facilitate the prognosis, diagnosis and monitoring of ocular surface disease in patients. It should also aid in the development of new therapeutics. Several individual biomarkers have been proposed, including lysozyme and lactoferrin [159], and MMP-9 [421]. Grouping biomarkers into an optimized panel may offer superior sensitivity and sensitivity (Table 8) [407,409,410,421,442,450,493], as revealed by Receiver Operating Characteristic (ROC) curve analysis which plots true positive (sensitivity) and false positive (specificity) rates respectively on y-and x-axes. The Area Under the upward rising Curve (AUC) is a measure of test accuracy, with a perfect test having an AUC value of 1, but is surprisingly lacking for several assayed biomarkers (Table 8). However, all these markers require further validation in independent studies and cohorts of participants from different geographic populations, age and etiology of DED. Panels in their present form require mass spectrometry, an approach currently not appropriate for day-to-day patient analysis. Evolution of panels into an integrated immunological assay holds considerable promise.

Table 8.

Diagnostic power of individual and panels of biomarkers for dry eye discrimination relative to normal controls.

Biomarker (reference) Dry Eye Number of subjects (dry-eye/normal) AUC CA (%) Sensitivity (%) Specificity (%) PPV NPV Technique Reference
Individual
ALB ADDE 160/45 0.729 42.5 90.9 97.1 Protein LabChip Bionalyzer Versura et al., 2013 [450]
ANXA1 ADDE 63/43 89.7 92.0 85.6 2D-PAGE/MALDI-TOF MS Soria et al., 2013 [409]
ANXA11 ADDE 63/43 76.0 84.0 61.6 2D-PAGE/MALDI-TOF MS Soria et al., 2013 [409]
AZGP1 ADDE 160/45 0.759 56.9 87.5 81.8 Protein LabChip Bionalyzer Versura et al., 2013 [450]
CST4 ADDE 63/43 67.7 90.2 27.2 2D-PAGE/MALDI-TOF MS Soria et al., 2013 [409]
ENO1 ADDE 56/40 0.850 77.0 88.5 ITRAQ Zhou et al., 2009 [407]
LCN1 ADDE 160/45 0.843 65.5 94.9 97.1 Protein LabChip Bionalyzer Versura et al., 2013 [450]
LTF ADDE 160/45 0.812 44.2 93.2 94.5 Protein LabChip Bionalyzer Versura et al., 2013 [450]
LYZ ADDE 160/45 0.826 57.8 96.6 97.9 Protein LabChip Bionalyzer Versura et al., 2013 [450]
MMP9 ADDE 143/63 85.0 94.0 97.0 73.0 Immunocapture Sambursky et al., 2013 [493]
ADDE 146/91 81.0 98.0 Immunocapture Sambursky et al., 2014 [497]
PLDE 14/34 57.1 Immunocapture Chan et al., 2015 [421]
PIP ADDE 56/40 0.810 71.7 87.1 ITRAQ Zhou et al., 2009 [407]
PLAA ADDE 63/43 92.3 94.2 88.0 2D-PAGE/MALDI-TOF MS Soria et al., 2013 [409]
PRDX5 ADDE 63/43 75.1 91.6 45.6 2D-PAGE/MALDI-TOF MS Soria et al., 2013 [409]
S100A6 ADDE 63/43 90.6 93.8 84.8 2D-PAGE/MALDI-TOF MS Soria et al., 2013 [409]
TF ADDE 160/45 0.885 70.8 96.7 98.1 Protein LabChip Bionalyzer Versura et al., 2013 [450]
Panel
PRR4, S100A8, SERPINA1, SMR3B plus two unknowns (3700 and 3916 Da) ADDE 88/71 0.93 90.0 90.0 SELDI-TOF-MS Grus et al., 2005 [442]
ENO1, PIP, LCN1, S100A9 ADDE 56/40 0.96 91.0 90.0 iTRAQ Zhou et al., 2009 [407]
ANXA1, ANXA11, CST4, PLAA, S100A6 ADDE 63/43 1.00 94.6 96.0 92.0 2D-PAGE/MALDI-TOF MS Soria et al., 2013 [409]
B2M, PRR4 (4043 & 4078 Da) S100A8, SCGB1D1, SCGB2A1 ADDE 40/39 1.00 100.0 100.0 SELDI-TOF-MS Boehm et al., 2013 [410]
MGD 40/39 1.00 100.0 100.0

ALB = albumin; ANXA1 = annexin A1; ANXA11 = annexin A11; AZGP1 = alpha-2-glycoprotein 1; CST4 = cystatin S; ENO1 = Enolase 1; LCN1 = Lipocalin 1; LTF = lactoferrin; LYZ = lysozyme; MMP9 = matrix metalloproteinase 9; PIP = prolactin-induced protein; PLAA = phospholipase A2 activating protein; PRDX5 = peroxiredoxin 5; S100A6 = S100 calcium-binding protein A6; TF = transferrin; PRR4 = proline-rich protein 4; S100A8 = S100 calcium-binding protein A8; SERPINA1 = serpin family A member 1; S100A9 = S100 calcium-binding protein A9; B2M = beta-2 microglobulin; SCGB1D1 = secretoglobin family 1D member 1; SCGB2A1 = secretoglobin family 2A member 1. ADDE = aqueous deficient dry eye; MGD - meibomian gland dysfunction.

Biomarkers should help identify and screen out individuals at risk of developing DED prior to refractive eye surgery, that itself can promote extended dryness. To date, biomarker analysis has been carried out after LASIK [421,494] but not before. Prophylactic treatment for at risk individuals might also be possible, for example treatments that promote natural basal tearing, ocular surface homeostasis and corneal sensory innervation. Similarly, diagnosis of the type [409] and severity [407,410,430,432,444,450] of dry eye, or other ocular surface pathology [446], should become much more precise. One can also imagine a precision medicine approach to dry eye, whereby deficiency of one or several functionally-relevant tear proteins is topically corrected. Follow up of patients may also be more accurate, such as monitoring of tear MMP-9 after surgery for conjunctivochalasis [418], or for monitoring the increase of tear lacritin, lysozyme and prolactin-inducible protein, and decrease of enolase 1 (alpha), S100A6 and S100A11 coincident with reduced ocular surface inflammation one year after switching patients from preserved latanoprost to preservative-free tafluprost for glaucoma [495].

Currently, the only commercial options to investigate tear film biomarkers are Inflammadry® (MMP-9) and TearScan™ Lactoferrin Diagnostic Test Kit. Inflammadry detects ~40 ng/ml or more MMP-9, with indicated sensitivity and specificity values respectively of 85 and 94%, and negative and positive predictive values respectively of 73% and 97% [493]. Clinical studies suggest specificity and sensitivity of 80% in patients with pseudoexfoliation syndrome [496], positive agreement of 81% and a negative agreement of 98% in dry eye patients [497], and 57% detection of MMP-9 in post-LASIK dry eye patients [421]. However, in a population with no previous diagnosis of DED and who were classified as having mild dry eye in the study, the Inflammadry assay was positive for only 11% of subjects with symptoms of dry eye [498], and in a study of symptomatic (DEQ-5 ≥6) subjects, only 39% tested positive with the Inflammadry test [499]. These data suggest that the Inflammadry test is better at detecting moderate to severe dry eye subjects. The lactoferrin assay has indicated sensitivity and specificity values, respectively, of 83 and 98%, and precision of <9% coefficient of variation. Both are semi-quantitative assays.

Proteomics and biomarker cell biology are progressively transforming ocular surface clinical research. From proteomics, new immunoassays will enhance prognosis and diagnosis. From cell biology, new therapeutics should emerge. Proteomic discovery will continue to be challenged by the nine orders of magnitude between the most (mg/mL) and least (pg/mL) abundant proteins [500]. Development of multiplexed systems that allow simultaneous quantification of several biomarkers in a single tear sample (commercially available only thus far for cytokines, interleukins and metalloproteinases - although development work is underway) [446,501503], and coordinated clinical validation studies are needed. More information is needed on biomarker dynamic range, cut-off definition, specificity, sensitivity, positive predictive value and negative predictive value, and linkage of biomarker detection to treatment.

Lagging behind immunoassay and discovery is an appreciation for the biology of biomarkers. Are there disease modifications, such as glycosylation, methylation, acetylation or covalent cross-linking, and disease-upregulated slice variants? N-glycosylation in climatic droplet keratopathy was the subject of one MS/MS study [445] and Perumal et al. [504] discovered different methylated isoforms of proline-rich protein 4 in tears for which disease or functional implications are unknown. Cross-linking by tissue transglutaminase (TGM2) in tears may play an important disease role, as conjunctival expression of TGM2 is elevated both in Sjögren syndrome and MGD [505], and substrates include S100A7, annexin A1, alpha-2-HS-glycoprotein, enolase 1, actin, beta-2-microglobulin, plasminogen [506], and lacritin [507]. Notably, TGM2 constitutively inactivates lacritin’s C-terminal ‘business end’ necessary for cell targeting [507], and more higher order protein complexes have been noted in dry eye [508], the latter also perhaps from mis-folding as a potential new direction in dry eye research.

4.4. Lipid-protein-mucin interactions

Secreted mucins may have an active role in lipid organization at the surface of the tear film [296,509] and phospholipid transfer protein functions to scavenge lipophilic substances from ocular mucins [510], but data supporting a direct interaction between mucins and lipids are lacking. In the epithelial glycocalyx, the transmembrane mucins MUC1 and MUC16 bind galectin-3 to promote ocular surface barrier function [382]. This interaction is dependent on the surface recognition carbohydrate residues and the capacity of galectin-3 to form surface lattices [511]. The altered mucin glycosylation observed in DED could lead to loss of galectin-3 binding affinity and release of cellular galectin-3 into the tear film. This hypothesis is supported by increased shed galectin-3 levels in dry eye tears concomitant with a galectin-3 proteolytic cleavage product with the potential to impair proper barrier function [383]. Clusterin, a cytoprotective chaperone, inhibits MMP-9 [512] and interacts with galectin-3 to ameliorate ocular surface barrier disruption by a sealing mechanism on the ocular surface [513]. Elevated levels of MMP-9 have been linked to epithelial barrier disruption.

The dominant lipid binding protein in tears is lipocalin-1 [255,514,515]. Lipocalin-1 (previously known as tear prealbumin) can bind a broad range of lipid ligands such as fatty acids, cholesterol and phosphatidyl lipids, but has a low affinity for lysophospholipids [515,516]. Lipocalin stabilizes the tear film by both scavenging lipids from the hydrophobic ocular surface into the mucoaqueous layer [517], and by interacting with tear polar lipids to reduce surface tension [286]. Whether lipocalin-1 is altered in ADDE is uncertain (Table 6). It is increased in tears of intolerant contact lens wearers [518], but decreased in tears of patients with MGD [468] and Sjögren syndrome [519] (Table 6).

Other proteins known to interact with lipids in the tear film include the tear surfactant proteins B and C [520], apolipoprotein D [521], lysozyme [299], keratins [300], secretoglobins (also known as lipophilins, mammaglobins, lacryglobin and uteroglobins) [522524] and phospholipid transfer protein [525]. Of these, only levels of lysozyme are altered in dry eye (down in ADDE and Sjögren syndrome dry eye, and in blepharitis – see Table 6).

Some soluble tear proteins interact with the epithelial cell surface to help promote ocular surface health. Lacritin is a multi-functional tear protein that in its monomeric form [507] binds cell surface heparan sulfate proteoglycan syndecan-1 [526] to promote basal tear secretion [438]. Binding of lacritin to syndecan-1 is dependent on heparanase modification of heparan sulfate to expose both a 3-O-sulfate site in cleaved heparan sulfate and the core protein sequence GAGAL [527]. Tear levels of active lacritin monomer are reduced in all forms of dry eye (Table 6) and in blepharitis [443]. Further, the levels of heparan sulfotransferases are downregulated in conjunctival cells from dry eye patients [376], which could influence heparanase cleavage and/or lacritin binding of syndecan-1 [527]. Vitelline membrane outer layer 1 homolog has been shown to interact with lysozyme and stabilize the tear film [492], but the relevance of this interaction to DED has not been yet established.

4.5. The tear metabolome

The aim of metabolomics is to identify, determine, interpret, and quantify the complex time-related concentration, activity, and flux of endogenous metabolites in cells, tissues, and other biosamples [528]. The mucoaqueous layer of the tear film contains electrolytes (such as sodium, potassium, calcium, magnesium, chlorine, phosphate, bicarbonate), protein fragments as peptides, and metabolites (such as amino acids, urea, glucose, lactate). Because the levels of the latter are low (confounded by low volumes of sample), the tear metabolome is poorly characterized [529].

The main analytical techniques implemented in metabolomics analysis are proton nuclear magnetic resonance spectroscopy (1H NMR) [530,531], MS, gas chromatography, and liquid chromatography-MS (LC-MS) [532,533]. Following information-dependent acquisition, directed liquid chromatography/tandem mass spectrometry (LC-MS/MS) coupled to peak mining can be performed, using isotope pattern matching [534]. NMR and GC-MS (gas chromatography mass spectrometry) offer high reproducibility and straightforward identification of metabolites. NMR is nondestructive, requires minimal sample preparation and offers the benefit of obtaining quantitative data simultaneously. However, it is the least sensitive of the above techniques and may be biased toward the detection of high-abundance metabolites.

The analysis includes extraction of metabolites, chromatographic separation and acquisition of data. Post-acquisition data processing typically comprises peak detection and alignment, followed by chemometric analysis to identify the marker ions. Metabolites are then identified through database searches, and their identities are confirmed by comparison of retention time and tandem mass spectra with pure standards [535]. However, no analytical platform can resolve and detect all the metabolites in a sample [534].

The study of tear metabolite profiling is of particular interest, as metabolomics provides an insight into the dynamic biochemical processes occurring in biological systems [536]. The most comprehensive description of human tear metabolome has been published by Chen et al., [534] who identified 60 metabolites representing diverse compound classes, of which 16 had been previously reported. To date, 85 metabolites have been identified in the tears of normal human subjects, 41 previously reported [537546] and 44 newly reported by Chen et al. [534].

Healthy and dry eye tears display differences. Reduced levels of carnitine in dry eye patients appear to associate with low retention of water in the mucoaqueous layer. This may contribute to hyper-evaporation associated with hyperosmolarity [542]. Nakatsukasa et al. [537] have identified 23 amino acids in healthy subjects and found reduced levels of taurine, methionine, and arginine in the severe dry eye patients, in comparison with those in healthy individuals, relative to a change to total amino acid concentration. The concentration of diadenosine polyphosphates increases in tears from dry eye subjects and those with Sjögren syndrome [547,548].

Most recently, differences between the metabolomic profiles of reflex tears in patients with mild-to-moderate and moderate DED and healthy subjects have been analyzed using 1H NMR spectroscopy [530]. Differences between tear metabolite levels were found for cholesterol, N-acetylglucosamine, glutamate, creatine, amino-n-butyrate, choline, acetylcholine, arginine, phosphoethanolamine, glucose, and phenylalanine (Table 9). The healthy subjects had elevated concentration of glycoproteins, lipids, cholesterol, leucine, glycerol, and glutamate in comparison with dry eye patients. In the dry eye samples, an increase in glucose and lactate and a decrease in formic acid and N-acetylglucosamine levels were found. Increased levels of serotonin have been correlated with dry eye signs and symptoms and higher concentrations of this neuro-transmitter have been found in symptomatic ADDE tears compared to tears of symptomatic dry eye patients with normal tear production [549].

Table 9.

Metabolite differences in the tears of dry eye tears in comparison to control subjects.

Metabolites Change in dry eye tears Reference
Cholesterol/lipids Galbis-Estrada et al., 2014 [530]
N-Acetylglucosamine Galbis-Estrada et al., 2014 [530]
Glutamate Galbis-Estrada et al., 2014 [530]
Carnitine Galbis-Estrada et al., 2014 [542]
Taurine Nakatsukasa et al., 2011 [537]
Methionine Nakatsukasa et al., 2011 [537]
Arginine Nakatsukasa et al., 2011 [537]
Serotonin Chhadva et al., 2015 [549]
Total creatine Galbis-Estrada et al., 2014 [530]
Arginine + phosphoetanolamine Galbis-Estrada et al., 2014 [530]
Phenylalanine Galbis-Estrada et al., 2014 [530]
Choline/acetylcholine Galbis-Estrada et al., 2014 [530]
Choline Galbis-Estrada et al., 2014 [530]
Glucose Galbis-Estrada et al., 2014 [530]
Amino-n-butyrate Galbis-Estrada et al., 2014 [530]
Diadenosine polyphosphates (Ap4A and Ap5A) Carracedo et al., 2010 [547] Peral et al., 2006 [548]

Finally, the effect of antioxidants and polyunsaturated fatty acid oral treatments on the profile of reflex tear metabolites has been demonstrated [531]. The metabolomics methods have some limitations for the study of human tears and further studies are required to broaden our knowledge of the tear metabolome. However, it is clear that the metabolome reflects the status of local and systemic biological processes, opening new routes of identification of potential biomarkers for disease diagnosis and treatment monitoring.

5. Translational dry eye animal models of tear film

There are several genetic, surgically or dry chamber/drug-induced models of DED in different species [489,550554]. The most characterized model is the desiccating stress murine model without or with prior injection of scopolamine to block parasympathetic stimulation of tear secreting machinery [555574]. In this dry eye mouse desiccating stress model, the levels of several tear cytokines and chemokines and some others molecules increase as a result of dry eye induction (Table 10). These changes are similar to those found in human tears (although not always evaluated with the same techniques; see Table 6) [407,409411,415,424,425,427,429,432,442,444,446,447,450]. Some, such as IL-17, do not always behave similarly [429]. The regulatory effects of cytokines and chemokines in tears have been observed after depletion of antigen presenting cells in mice [566]. It should be noted that the most commonly expressed mRNAs in mouse and human lacrimal gland differ substantially [575].

Table 10.

Animal tear proteins that appear to decrease or increase in induced dry eye.

Dry Eye Animal Model Tear Change from Normal & Induced DE Technology for Tear Analysis Reference
Aqueous deficient dry eye (lacrimal insufficiency) Rabbit Induced autoimmune dacryoadenitis (IAD) ↑S100A6, ↑S100A9, ↑ALB, ↓TF, ↓PIP, ↓pIgR, ↓Ig ƴ chain C region iTRAQ + nanoLC-nanoESI-MS/MS Zhou et al., 2013 [576]
Combined lacrimal insufficiency and evaporative dry eye models Mice Desiccating stress (DS) model and scopolamine ↑IL-17 Immunobead assay De Paiva et al., 2009 [555], Dursun et al., 2002 [568]
↑IL-1β, TNF-α, IL-2, IL-6, CCL5, CXCL10, IFN-ƴ, IL-17, IL-10 Luminex Schaumburg et al., 2011 [566]
↑IL-1α, TNF-α, IL-2, IL-12, IFN-γ, IL-4 and IL-10 (mouse strain dependent) Immunobead assay Corrales et al., 2007 [574]
↑IL-1β, MMP9 ELISA, Zymography Luo et al., 2004 [591]
↑ SP-D Western immunoblot Heimer et al., 2013 [592]
↑ MMP1, MMP3, MMP9 Activity assay Corrales et al., 2006 [573]
Aqueous deficient Mice IL-2r alpha (CD25) knockout mice (develop Sjögren syndrome-like disease) ↑ IL-1 and IFN-γ Luminex De Paiva et al., 2010 [554]
Preganglionic parasympathetic denervation Rabbit Parasympathetic denervation ↑CGRP after cyclosporine treatment ELISA Toshida et al., 2009 [593], Toshida et al., 2007 [594]

S100A6 = S100 calcium-binding protein A6; S100A9 = S100 calcium-binding protein A9; ALB = albumin; TF = transferrin; PIP = prolactin-induced protein; pIgR = polymeric immunoglobulin receptor; Ig = immunoglobulin; IL = interleukin; TNF = tumor necrosis factor; CCL = C-C motif chemokine ligand; CXCL = C-X-C motif chemokine ligand; IFN = interferon; MMP-9 = matrix metallopeptidase 9; SP-D = surfactant protein-D; CGRP = calcitonin gene-related peptide.

Zhou et al. [576] have reported on similarities to human dry eye tear protein expression in a rabbit model of Sjögren syndrome-associated dry eye induced autoimmune dacryoadenitis (Table 10). The expression of S100A6, S100A9, and serum albumin was upregulated in this model, whereas TF, PIP, polymeric immunoglobulin receptor (pIgR) and Ig gamma chain C region were downregulated. In this animal model, the changes in the expression of S100A6, TF, PIP, and pIgR were verified. The protein pIgR may be unique to Sjögren syndrome, while the remaining proteins were also found in the tears of non-Sjögren syndrome dry eye patients [407,409,448].

The Harderian gland, that releases secretions onto the ocular surface and fills the orbit of all animals with the exception of primates, is a potential complicating factor that is largely often overlooked. Butovich et al. [577] have compared the mouse, canine, rabbit, and human meibum lipidomes using HPLC and GLC. This study demonstrated that mouse and human meibomian lipidomes are most closely matched, and the canine lipidome is the second closest. Thus, the rabbit is not the best model of the human tear film for studying the lipidome biochemistry and biophysics [577].

Some progress has also been made in the search for appropriate dry eye models with respect to severity. Three rabbit dry eye models for mild (lacrimal gland and nictitating membrane removal), moderate (conjunctival burn with 50% trichloroacetic acid), and severe manifestations of dry eye (combination of both methods), have been developed [578]. However, there is a lack of studies analyzing tear instability in these models [579].

6. Non-pathophysiological factors impacting the tear film

6.1. Contact lenses

Contact lens wear increases the risk of developing dry eye from between 2.01 and 2.96 times [580582]. Contact lenses compartmentalize the tear film into two layers; the outer pre-lens tear film layer in front of the lens, and the inner post-lens tear film layer, which lies between the lens and the cornea [583,584]. With lens wear, the tear film undergoes extensive biophysical and biochemical changes, which have the potential to influence tear function and/or contact lens tolerance.

In 2013 TFOS published the outcomes of a workshop on Contact Lens Discomfort, including a subcommittee report of Contact Lens Interactions With the Tear Film [54]. The following were associated with worsening of comfort during lens wear: increases in the rapidity of tear film breakup time on the cornea or over the surface of a contact lens; higher contact lens water content; the grade of tear ferning (grade 0 vs. grades 1 or 2); increased tear evaporation rate during contact lens wear; reduced tear flow rate and tear volume; impaired tear film lipid layer; mucin degradation on contact lenses; increases in the concentrations of lipocalin-1, secretory phospholipase A2 (and activity), and lipid aldehyde products. The following section will concentrate on references published since that report.

In cross-sectional studies there are no statistically significant differences in tear film osmolarity between contact lens wearers and non-lens wearers [131,585]. The concentration of phospholipase A2 levels in tears has been associated with an increased concentration of malondialdehyde, which is also a by-product of phospholipid degradation and shorter TBUT over a contact lens [586].

The limited thickness of the post-lens tear film (1–2 μm) has clinical implications, as the stagnation of tears beneath a contact lens is believed to be one of the major causes of complications seen with contact lens wear, including mucin ball formation and ocular surface staining [587]. Technological limitations have not allowed for the accurate and non-invasive determination of fluid circulation between the pre- and post-lens tear film, a phenomenon commonly referred to as tear exchange or tear turnover. Fluorophotometry studies on tear exchange have demonstrated that tear exchange occurs in 30 min with soft lenses [588,589], which is longer than that with rigid lenses or no lens wear [590]. This large difference probably relates to lens size and material stiffness, and therefore lens movement with the blink. Silicone hydrogel materials were reported to improve tear exchange by 5% compared to hydrogels [589] due to the slightly higher modulus, which improves lens movement. Good tear exchange may minimize the risk of adverse sequelae via timely flushing of debris and metabolic byproducts from the ocular surface. Innovations in soft lens designs, in an attempt to increase tear exchange, have not proven clinically effective thus far [587].

A thicker lipid layer during contact lens wear is associated with increased stability of the tear film, increased interblink period and decreased tear evaporation rate [586]. Using interferometry, the lipid layer thickness over soft contact lenses was estimated to be around 15 nm [586]. Reduced pre-lens TBUT values have been associated with increased symptoms of discomfort in hydrogel contact lens wearers [586]. A prospective, randomized, crossover study, showed that a daily disposable silicone hydrogel contact lens manufactured from delefilcon A permitted significantly longer NIBUT compared to lenses made from narafilcon A and filcon II-3, in addition to inducing significantly less corneal staining after 16 h of wear [595]. However, the study found no improvement in comfort with any particular contact lens [595]. Use of a liposomal eye spray has been reported to improve pre-lens TBUT [586]. Recently, a prospective, randomized, double blind study reported a favorable effect of Omega 3 fatty acid supplements in improving tear film stability, tear film breakup time and wear comfort in contact lens wearers [596].

The hypothesis that inflammation may be involved in contact lens discomfort has not been supported by two studies examining levels of cytokines in tears of contact lens wearers. One study examining tear cytokine concentrations of symptomatic and asymptomatic lens wearers found no difference in the concentration of 11 cytokines in tears, including IL-1β, IL-6 and IL-8 [597]. The other study found no change in the concentration of 14 cytokines in tears over the course of a day wearing contact lenses, even though comfort was significantly reduced [440]. The latter study did find the concentration of VEGF was correlated with ocular comfort, but the change in VEGF concentration was more pronounced when contact lenses were not worn rather than during lens wear.

Although previous studies evaluating transmembrane mucins during contact lens wear reported conflicting results, using fluorescein-labeled wheat germ agglutinin (F-WGA) as a marker, Fukui et al. [598] reported less fluorescence intensity in soft contact lens wearers. Observing a positive correlation between tear film breakup time and F-WGA fluorescence, the authors suggested that a decrease or compositional alteration of ocular surface glycocalyx could be a factor in contact lens-induced dryness. Disturbance of the glycocalyx during contact lens wear may be due to slower epithelial turnover and consequently decreased quality of surface mucin, contact lens-induced friction and resultant inflammation, or restricted tear exchange resulting in restricted delivery of secreted mucins on the corneal surface [598].

What is still unclear is the natural history of contact lens discomfort. Are people who experience contact lens discomfort normal before lens wear and at risk of developing symptoms of dry eye even on contact lens removal entirely as a result of wearing lenses? Or are those people predisposed to dry eye (perhaps having subclinical signs or symptoms) and contact lens wear tips them towards dry eye?

6.2. Hormonal factors

Low androgen levels have been reported to be consistently associated with dry eye [599]. Biophysical and biochemical changes in the tear film in postmenopausal women have been investigated in a few studies. Tear osmolarity in postmenopausal women with dry eye symptoms were shown to be significantly higher than those without symptoms [124], and sex hormone levels correlated with tear osmolarity [600,601]. Tear ferning tests were also reported to have higher grades in postmenopausal women with dryness complaints than asymptomatic postmenopausal women [124]. As for the biochemical changes, although tear film levels of lipocalin and lysozyme do not seem to change in postmenopausal women with dry eye [602], in one study lactoferrin levels were reported to significantly decrease [603]. The effect of hormone replacement therapy is controversial; some studies suggested it improves tear function tests and increases goblet cell density [604,605], whereas, others reported hormone replacement therapy (particularly estrogen alone) to be associated with a higher risk of dry eye [606]. A further, more exhaustive discussion on this topic is provided in the TFOS DEWS II Sex, Gender, and Hormones report [607].

Thyroid eye disease may be associated with tear function abnormalities. Increased dryness symptoms, decreased tear film breakup time, decreased corneal sensitivity, and increased ocular surface staining have been reported both during the active and inactive phases of Graves’ ophthalmopathy [608,609]. Increased tear osmolarity [610,611] and increased levels of inflammatory mediators such as IL-1β, IL-6, IL-8 and plasminogen activator inhibitor-1 have also been found in the tear film of patients with active Graves’ ophthalmopathy [609,612].

6.3. Environmental factors

The use of visual display units (VDU) is also a factor that is associated with dryness symptoms, even in young subjects. Up to 27.4% of VDU users reportedly have high tear osmolarity levels [613]. Kojima et al. [614] reported that long-term VDU users (using digital devices more than 4 h per day) exhibited a significantly reduced TMH and increased symptom scores compared to short-term VDU users and controls. VDU-related dryness has been attributed to reduction in spontaneous blink rate [615618], predominance of incomplete blinks [615], the presence of MGD [619], lipid layer instability and decreased TBUT [614,615]. In normal subjects, NIBUT has been shown to decrease significantly after 30 min of playing a computer game [93]. The resultant tear film hyperosmolarity was suggested to be reliable in diagnosing dry eye in VDU users, and osmolarity levels were shown to correlate with tear function tests and tear film breakup levels [620]. Frequent breaks from VDU use, blinking awareness training and moist cool air devices were reported to provide relief of ocular dryness and improvement of tear stability for these subjects [238,621].

In vitro and in vivo studies have shown that increasing the ambient temperature increases the tear evaporation rate, TBUT, lipid layer thickness, and ocular surface temperature [254,622]. High altitude exposure, with cold and dry air, leads to a decrease in TBUT and an increase in tear film osmolarity [148]. Controlled environmental chamber studies provide a way of controlling the environmental temperature and humidity independently or together. One controlled environmental chamber study showed that, after 2 h exposure in the chamber, TBUT decreased, corneal staining and tear MMP-9 levels increased in both symptomatic (mild to moderate dry eye) and asymptomatic control subjects exposed to 5% (low) humidity in a desiccating environment [419]. The authors concluded that acute adverse changes in environmental conditions could lead to significant functional deterioration in subjects with or without dry eye symptoms [419]. In another study, changing the humidity from 40% (normal) to 5% (low) significantly increased tear evaporation rate, decreased lipid layer thickness, increased tear instability, and reduced tear production [251]. Exposing non-adapted contact lens wearers to a controlled environmental chamber for 20 min increased tear osmolarity, evaporation rates, visual symptoms and tear instability significantly for hydrogel wearers compared to silicone hydrogel wearers [234]. Another factor that may impact the tear film, leading to signs and symptoms of dry-eye, is air pollution [623]. Exposure to traffic-derived air pollution reduces the goblet cell numbers in the conjunctiva [624]. Further details can be found in the TFOS DEWS II Iatrogenic Dry Eye report [625].

In conclusion, several factors, including contact lens wear and environmental conditions (such as VDU use, ambient temperature and humidity) significantly affect tear film stability and function. However, it is yet to be established if any of these changes can lead to the development of chronic DED.

7. Conclusions and new frontiers

A two layered tear film model of the tear film’s structure is preferred over the traditional three layered model, with a mucoaqueous gel layer lying beneath, but at least partly integrated with, an overlying lipid layer. The mucoaqueous layer makes up the bulk of the thin (2–6 μm) tear film, but most data on thickness is from the precorneal tear film and there is virtually no data on the pre-conjunctival and other compartment tear films. The lipid layer is very thin, at approximately 42 nm.

Tears are distributed over the ocular surface by the combined actions of the lids, capillary movement and upward drift of the lipid layer. Tear film thinning and breakup occurs mainly as a result of evaporation of water from the tear film, and breakup times are usually twice as rapid as normal, for people with DED. The development of lateral shearing interferometry or double pass methods may improve discrimination of TBUT between normal and dry eye subjects.

It is important that future studies be clear about how, and from which compartment, tears are collected, and the type of tears (reflex, basal, or after sleep). Most studies that have measured tear film osmolarity have used tears from the lower tear film meniscus and have shown increases in tear osmolarity with dry eye. However, it is likely that the osmolarity of the lower meniscus does not accurately reflect the osmolarity over the rest of the ocular surface. Development of techniques to measure tear film osmolarity over the cornea, ideally before and after the tear film ruptures, may provide important insight into diagnosis, improve discrimination of the type of DED present, assist with development of new treatments, and provide more accurate tests for monitoring treatments for DED. Tear evaporation rates increase with dry eye by approximately twofold. However, the lack of commercially available, reliable evaporimeters to measure this in the clinic limits its use to specialized facilities. The development of evaporimeters that can be used in clinical settings and take into account variations in ambient humidity and temperature would be extremely valuable.

Mucins help to stabilize and spread tears by binding, via their high levels of glycosylation, to water. The lipid layer reduces the surface tension of tears, allowing the thin film to avoid collapse. The lipid layer is probably responsible for reducing evaporation of tears, but how this occurs and the role of proteins and mucins in this process requires further investigation. Also, further investigation is needed to understand how proteins and mucins form the mucoaqueous gel.

The traditional belief that the tear film lipid layer can reduce water evaporation from the mucoaqueous layer has been recently questioned, by experiments in vitro that have shown that meibum or artificial lipid mixtures can only reduce evaporation of underlying water by ≤ 8%. It may be that intercalating proteins and mucins in the tear film lipid layer are needed to reduce evaporation. However, if this is the case, then results showing the importance of the ability of the tear film lipid layer to reduce the surface tension of tears and that the presence of proteins in this layer disrupts this system are difficult to reconcile. Also, proteins in meibum impair its spreading and worsen its viscoelasticity. Clearly, more studies in these areas are needed.

The tear lipidome has been explored in great detail since the publication of the first TFOS DEWS report in 2007. At least five, probably more, lipid classes are present, with each class having many different types due to the large number of possible fatty acids than can be esterified within each class. The majority of lipids in meibum and the tear film lipidome are wax and cholesterol esters. Debates regarding the presence of phospholipids in meibum have largely been settled and experiments show very low levels of this lipid class in meibum. However, the presence of relatively high levels of phospholipids in the tear film lipidome then begs the questions of where these originate from, and indeed if they are really in the tear film lipid layer or are extracted from lipid-binding proteins in tears such as lipocalin (and thereby sequestered away from the tear film lipid layer). Studies have shown that dry eye may be related to changes in the tear film lipidome, especially to changes in the concentration of OAHFAs, but more work is required to confirm these findings. Multi-laboratory and multi-national testing of common pooled meibum and tear lipid extracts is required to provide baseline lipidome data that can then be used as a gold standard to examine the effects of dry eye and its treatments.

There are various mucins found in tears, with the major soluble mucin being MUC5AC and the transmembrane mucins being MUC1, MUC4 and MUC16. However, mRNA for other mucins has been found in ocular surface epithelial cells, and so may be present in tears, although perhaps in low concentrations. There are changes to the concentration of MUC5AC, MUC1, MUC4 and MUC16 in tears of dry eye subjects, but not all studies agree on the magnitude or direction of the changes. However, decreased MUC5AC expression and alteration in mucin glycosylation in dry eye is a consistent finding among most studies. Further studies using the latest glycomic, proteomic and genomic techniques will be beneficial in helping to determine the role of changes to mucin in DED.

There has been an explosion in the number of proteins known to be present in the tear film proteome, with approximately 1800 proteins being present. The use of non-biased MS methods to evaluate changes to proteins, along with more traditional immunoassays, has helped to show that several proteins change from normal levels in dry eye tears. These proteins can be used as biomarkers, which might help to predict, diagnose and even, in some cases, be therapeutic for dry eye. However, MS is not appropriate for converting into point-of-care analysis, whereas immunological techniques may be. To date, many candidate biomarkers found by MS have yet to be validated by immunological techniques. Furthermore, validating the biomarkers in separate populations is often lacking, resulting in a lack of data on sensitivity and specificity. Data that have been generated on specificity and sensitivity of protein biomarkers show that panels of at least four proteins are superior to individual proteins. More information is needed on biomarker dynamic range, cut-off definition, specificity, sensitivity, positive predictive value and negative predictive value, and linkage of biomarker detection to treatment. Also, testing application of biomarkers in animal models to demonstrate whether they can alleviate dry eye signs is lacking for many.

Investigations into the tear film metabolome have been small and sporadic, but much could be learned from more research in this area. The metabolome is a measure of endogenous metabolites in tears. Due to the fact that these metabolites can be many and varied, sophisticated laboratory techniques are needed during the discovery phase, and so development of assays appropriate for point-of-care use will need to be made for any metabolites that can be used as biomarkers. Changes to amino acids or their derivatives in the tear film of dry eye subjects have been found to be possible markers of the disease.

This subcommittee also examined the effect of non-pathophysiological factors impacting the tear film. Epidemiological evidence pointing to the high risk of having dry eye if the person has been a contact lens wearer has focused attention in this area, as had the publication of the TFOS report of Contact Lens Discomfort in 2013 [54]. Recent data does not support an underlying inflammatory response being associated with contact lens discomfort, but there are clearly changes to the tear lipid layer and mucins during contact lens wear. Whether these changes are transient or more permanent has not been investigated. Furthermore, it is unclear whether those people who have increased discomfort or dryness symptoms during contact lens wear were predisposed to becoming dry eye patients even if they did not wear lenses, or whether lens wear accelerated their path toward DED. With our increasing use of electronic means to communicate, the associated dry eye symptoms that can occur and how to manage or treat these is an area worthy of research in the future.

The tear film is clearly required for a healthy ocular surface. There are several changes to the tear film that occur in dry eye. How biochemical changes relate to biophysical changes, and how these relate to clinical manifestations of tear film stability and volume are not currently known, although progress is being made in many of these areas. The development of a holistic model of tear film structure and function, and changes that occur during dry eye will likely be forthcoming in response to current and future research.

Acknowledgments

F (Financial Support), I (Personal Financial Interest), E (Employment), C (Consultant), P (Patent), R (Recipient), N (No Commercial Relationship), S (non-remunerative).

Footnotes

Financial disclosures

M. Willcox, Chair: Alcon Laboratories, Allergan, Coopervision, Johnson & Johnson Vision Care, Ophtecs (F); Allergan, Minomic International Pty., Ophtecs, Warm Contacts (C); Allergan, Optecs (R); Minomic International Pty., Ophtecs (S)

P. Argüeso: Otsuka, Santen (F)

G. Gerogiev: Menicon Co., Ltd., Otsuka Pharmaceuticals Co., Ltd., Rohto Pharmaceutical, Co., Ltd., Santen Pharmaceutical Co., Ltd., (F)

J. Holopainen: None.

G. Laurie: TearSolutions, LLC (I); Ocular Tear Growth Factor-Like Protein US 7,320,870; 7,459,440; Japan 4288070; Canada 2,438,334; Use of Lacritin in Promoting Ocular Cell Survival US 7,648,964 B2; Eur 1766768; Japan 512278; Lacritin-Syndecan Interactions US 7,932,227 B1; Compositions and Methods for the Delivery of Drugs to the Ocular Surface by Contact Lenses PCT/US2014/32412; Compositions and Methods for Detecting, Diagnosing, Preventing and Treating Dry Eye PCT/US2015/19964 (P)

T. Millar: None.

E. Papas: None.

J. Rolland-Thompson: None.

T. Schmidt: Johnson & Jonson Vision Care, Contamac, Lubris (F); Lubris (I); Johnson & Johnson Vision Care, Allergan, Lubris (C); Lubris (P); Johnson & Johnson Vision Care (R)

U. Stahl: Advanced Vision Research, Alcon, Allergan, Contamac, CooperVision, Essilor, Inflamax, Johnson & Johnson Vision Care, Ocular Dynamics, Oculus, Safilens, TearLab, TearScience (F)

T. Suarez: Bioftalmik co-founder and investor (I), (E); Co-inventor of a patent related with dry eye biomarkers (EP2603797 European Patent & US20140057307A1 USA Patent) (P)

L. Subbaraman: Advanced Vision Research, Alcon, Allergan, Contamac, CooperVision, Essilor, Inflamax, Johnson & Johnson Vision Care, Ocular Dynamics, Oculus, Safilens, TearLab, TearScience (F); Johnson & Johnson (C), (R)

O. Ucakhan: None.

L. Jones (Harmonisation member): Advanced Vision Research, Alcon, Allergan, Contamac, CooperVision, Essilor, Inflamax, Johnson & Johnson Vision Care, Ocular Dynamics, Oculus, Safilens, TearLab, TearScience (F); Alcon, CooperVision, Johnson & Johnson Vision Care (C) (R)

References

  • 1.Wolff E. The muco-cutaneous junction of the lid margin and the distribution of the tear fluid. Trans Ophthalmol Soc U K. 1946;66:291–308. [Google Scholar]
  • 2.Holly FJ, Lemp MA. Tear physiology and dry eyes. Surv Ophthalmol. 1977;22:69–87. doi: 10.1016/0039-6257(77)90087-x. [DOI] [PubMed] [Google Scholar]
  • 3.Doane MG. Abnormalities of the structure of the superficial lipid layer on the in vivo dry-eye. In: Sullivan DA, editor. Lacrimal Gland, Tear Film, and Dry Eye Syndromes. USA: Springer; 1994. pp. 489–93. [DOI] [PubMed] [Google Scholar]
  • 4.Yokoi N, Bron AJ, Georgiev GA. The precorneal tear film as a fluid shell: the effect of blinking and saccades on tear film distribution and dynamics. Ocul Surf. 2014;12:252–66. doi: 10.1016/j.jtos.2014.01.006. [DOI] [PubMed] [Google Scholar]
  • 5.Wang J, Aquavella J, Palakuru J, Chung S, Feng C. Relationships between central tear film thickness and tear menisci of the upper and lower eyelids. Invest Ophthalmol Vis Sci. 2006;47:4349–55. doi: 10.1167/iovs.05-1654. [DOI] [PubMed] [Google Scholar]
  • 6.Savini G, Barboni P, Zanini M. Tear meniscus evaluation by optical coherence tomography. Ophthalmic Surg Lasers Imaging. 2006;37:112–8. [PubMed] [Google Scholar]
  • 7.Shen M, Li J, Wang J, Ma H, Cai C, Tao A, et al. Upper and lower tear menisci in the diagnosis of dry eye. Invest Ophthalmol Vis Sci. 2009;50:2722–6. doi: 10.1167/iovs.08-2704. [DOI] [PubMed] [Google Scholar]
  • 8.King-Smith PE, Fink BA, Hill RM, Koelling KW, Tiffany JM. The thickness af the tear film. Curr Eye Res. 2004;29:357–68. doi: 10.1080/02713680490516099. [DOI] [PubMed] [Google Scholar]
  • 9.King-Smith PE, Fink BA, Fogt N, Nichols KK, Hill RM, Wilson GS. The thickness of the human precorneal tear film: evidence from reflection spectra. Invest Ophthalmol Vis Sci. 2000;41:3348–59. [PubMed] [Google Scholar]
  • 10.Chen Q, Wang J, Tao A, Shen M, Jiao S, Lu F. Ultrahigh-resolution measurement by optical coherence tomography of dynamic tear film changes on contact lenses. Invest Ophthalmol Vis Sci. 2010;51:1988–93. doi: 10.1167/iovs.09-4389. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.King-Smith PE, Kimball SH, Nichols JJ. Tear film interferometry and corneal surface roughness. Invest Ophthalmol Vis Sci. 2014;55:2614–8. doi: 10.1167/iovs.14-14076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Holly FJ, Lemp MA. Surface chemistry of the tear film: implications for dry eye syndromes, contact lenses, and ophthalmic polymers. Cont Lens Soc Am J. 1971;5:12–9. [Google Scholar]
  • 13.Holly FJ. Formation and rupture of the tear film. Exp Eye Res. 1973;15:515–25. doi: 10.1016/0014-4835(73)90064-x. [DOI] [PubMed] [Google Scholar]
  • 14.Mantelli F, Mauris J, Argueso P. The ocular surface epithelial barrier and other mechanisms of mucosal protection: from allergy to infectious diseases. Curr Opin Allergy Clin Immunol. 2013;13:563–8. doi: 10.1097/ACI.0b013e3283645899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Govindarajan B, Gipson IK. Membrane-tethered mucins have multiple functions on the ocular surface. Exp Eye Res. 2010;90:655–63. doi: 10.1016/j.exer.2010.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Gipson IK, Spurr-Michaud S, Tisdale A. Human conjunctival goblet cells express the membrane associated mucin MUC16: localization to mucin granules. Exp Eye Res. 2015;145:230–4. doi: 10.1016/j.exer.2015.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Gipson IK, Argueso P. Role of mucins in the function of the corneal and conjunctival epithelia. Int Rev Cytol. 2003;231:1–49. doi: 10.1016/s0074-7696(03)31001-0. [DOI] [PubMed] [Google Scholar]
  • 18.Kautto L, Nguyen-Khuong T, Everest-Dass A, Leong A, Zhao Z, Willcox MD, et al. Glycan involvement in the adhesion of Pseudomonas aeruginosa to tears. Exp Eye Res. 2016;145:278–88. doi: 10.1016/j.exer.2016.01.013. [DOI] [PubMed] [Google Scholar]
  • 19.Zhou L, Zhao SZ, Koh SK, Chen L, Vaz C, Tanavde V, et al. In-depth analysis of the human tear proteome. J Proteomics. 2012;75:3877–85. doi: 10.1016/j.jprot.2012.04.053. [DOI] [PubMed] [Google Scholar]
  • 20.Azkargorta M, Soria J, Ojeda C, Guzman F, Acera A, Iloro I, et al. Human basal tear peptidome characterization by CID, HCD, and ETD followed by in silico and in vitro analyses for antimicrobial peptide adentification. J Proteome Res. 2015;14:2649–58. doi: 10.1021/acs.jproteome.5b00179. [DOI] [PubMed] [Google Scholar]
  • 21.Brown SH, Kunnen CM, Duchoslav E, Dolla NK, Kelso MJ, Papas EB, et al. A comparison of patient matched meibum and tear lipidomes. Invest Ophthalmol Vis Sci. 2013;54:7417–24. doi: 10.1167/iovs.13-12916. [DOI] [PubMed] [Google Scholar]
  • 22.McDonald JE, Brubaker S. Meniscus-induced thinning of tear films. Am J Ophthalmol. 1971;72:139–46. doi: 10.1016/0002-9394(71)91604-7. [DOI] [PubMed] [Google Scholar]
  • 23.Miller KL, Polse KA, Radke CJ. Black-line formation and the “perched” human tear film. Curr Eye Res. 2002;25:155–62. doi: 10.1076/ceyr.25.3.155.13478. [DOI] [PubMed] [Google Scholar]
  • 24.King-Smith PE, Hinel EA, Nichols JJ. Application of a novel interferometric method to investigate the relation between lipid layer thickness and tear film thinning. Invest Ophthalmol Vis Sci. 2010;51:2418–23. doi: 10.1167/iovs.09-4387. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Peng CC, Cerretani C, Li Y, Bowers S, Shahsavarani S, Lin MC, et al. Flow evaporimeter to assess evaporative resistance of human tear-film lipid layer. Ind Eng Chem Res. 2014;53:18130–9. [Google Scholar]
  • 26.Bron AJ, Tiffany JM, Gouveia SM, Yokoi N, Voon LW. Functional aspects of the tear film lipid layer. Exp Eye Res. 2004;78:347–60. doi: 10.1016/j.exer.2003.09.019. [DOI] [PubMed] [Google Scholar]
  • 27.Knop E, Knop N, Millar T, Obata H, Sullivan DA. The international workshop on meibomian gland dysfunction: report of the subcommittee on anatomy, physiology, and pathophysiology of the meibomian gland. Invest Ophthalmol Vis Sci. 2011;52:1938–78. doi: 10.1167/iovs.10-6997c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Rosenfeld L, Fuller GG. Consequences of interfacial viscoelasticity on thin film stability. Langmuir. 2012;28:14238–44. doi: 10.1021/la302731z. [DOI] [PubMed] [Google Scholar]
  • 29.Georgiev GA, Yokoi N, Ivanova S, Tonchev V, Nencheva Y, Krastev R. Surface relaxations as a tool to distinguish the dynamic interfacial properties of films formed by normal and diseased meibomian lipids. Soft Matter. 2014;10:5579–88. doi: 10.1039/c4sm00758a. [DOI] [PubMed] [Google Scholar]
  • 30.Svitova TF, Lin MC. Racial variations in interfacial behavior of lipids extracted from worn soft contact lenses. Optom Vis Sci. 2013;90:1361–9. doi: 10.1097/OPX.0000000000000098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Raju SR, Palaniappan CK, Ketelson HA, Davis JW, Millar TJ. Interfacial dilatational viscoelasticity of human meibomian lipid films. Curr Eye Res. 2013;38:817–24. doi: 10.3109/02713683.2013.782050. [DOI] [PubMed] [Google Scholar]
  • 32.Dilly PN. Structure and function of the tear film. Adv Exp Med Biol. 1994;350:239–47. doi: 10.1007/978-1-4615-2417-5_41. [DOI] [PubMed] [Google Scholar]
  • 33.Cher I. A new look at lubrication of the ocular surface: fluid mechanics behind the blinking eyelids. Ocul Surf. 2008;6:79–86. doi: 10.1016/s1542-0124(12)70271-9. [DOI] [PubMed] [Google Scholar]
  • 34.Goto E, Tseng SC. Kinetic analysis of tear interference images in aqueous tear deficiency dry eye before and after punctal occlusion. Invest Ophthalmol Vis Sci. 2003;44:1897–905. doi: 10.1167/iovs.02-0818. [DOI] [PubMed] [Google Scholar]
  • 35.Yokoi N, Takehisa Y, Kinoshita S. Correlation of tear lipid layer interference patterns with the diagnosis and severity of dry eye. Am J Ophthalmol. 1996;122:818–24. doi: 10.1016/s0002-9394(14)70378-2. [DOI] [PubMed] [Google Scholar]
  • 36.Korb DR, Greiner JV. Increase in tear film lipid layer thickness following treatment of meibomian gland dysfunction. Adv Exp Med Biol. 1994;350:293–8. doi: 10.1007/978-1-4615-2417-5_50. [DOI] [PubMed] [Google Scholar]
  • 37.Goto E, Dogru M, Kojima T, Tsubota K. Computer-synthesis of an interference color chart of human tear lipid layer, by a colorimetric approach. Invest Ophthalmol Vis Sci. 2003;44:4693–7. doi: 10.1167/iovs.03-0260. [DOI] [PubMed] [Google Scholar]
  • 38.Wang JH, Fonn D, Simpson TL, Jones L. Precorneal and pre- and postlens tear film thickness measured indirectly with optical coherence tomography. Invest Ophthalmol Vis Sci. 2003;44:2524–8. doi: 10.1167/iovs.02-0731. [DOI] [PubMed] [Google Scholar]
  • 39.Schmoll T, Unterhuber A, Kolbitsch C, Le T, Stingl A, Leitgeb R. Precise thickness measurements of Bowman’s layer, epithelium, and tear film. Opt Vis Sci. 2012;89:E795–802. doi: 10.1097/OPX.0b013e3182504346. [DOI] [PubMed] [Google Scholar]
  • 40.Werkmeister RM, Alex A, Kaya S, Unterhuber A, Hofer B, Riedl J, et al. Measurement of tear film thickness using ultrahigh-resolution optical coherence tomography. Invest Ophthalmol Vis Sci. 2013;54:5578–83. doi: 10.1167/iovs.13-11920. [DOI] [PubMed] [Google Scholar]
  • 41.Huang JX, Tankam P, Aquavella JV, Hindman HB, Clarkson E, Kupinski M, et al. Tear film thickness estimation using optical coherence tomography and maximum-likelihood estimation. Invest Ophthalmol Vis Sci. 2015;56 E-Abstract 351. [Google Scholar]
  • 42.dos Santos VA, Schmetterer L, Groschl M, Garhofer G, Schmidl D, Kucera M, et al. In vivo tear film thickness measurement and tear film dynamics visualization using spectral domain optical coherence tomography. Opt Express. 2015;23:21043–63. doi: 10.1364/OE.23.021043. [DOI] [PubMed] [Google Scholar]
  • 43.Isreb MA, Greiner JV, Korb DR, Glonek T, Mody SS, Finnemore VM, et al. Correlation of lipid layer thickness measurements with fluorescein tear film break-up time and Schirmer’s test. Eye (Lond) 2003;17:79–83. doi: 10.1038/sj.eye.6700224. [DOI] [PubMed] [Google Scholar]
  • 44.Zhuang H, Zhou XT, Xu JJ. A novel method for pachymetry mapping of human precorneal tear film using Pentacam with fluorescein. Invest Ophthalmol Vis Sci. 2010;51:156–9. doi: 10.1167/iovs.08-3265. [DOI] [PubMed] [Google Scholar]
  • 45.King-Smith PE, Reuter KS, Braun RJ, Nichols JJ, Nichols KK. Tear film breakup and structure studied by simultaneous video recording of fluorescence and tear film lipid layer images. Invest Ophthalmol Vis Sci. 2013;54:4900–9. doi: 10.1167/iovs.13-11878. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Su TY, Chang SW, Yang CJ, Chiang HK. Direct observation and validation of fluorescein tear film break-up patterns by using a dual thermal-fluorescent imaging system. Biomed Opt Express. 2014;5:2614–9. doi: 10.1364/BOE.5.002614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Mishima S, Gasset A, Klyce SD, Jr, Baum JL. Determination of tear volume and tear flow. Invest Ophthalmol. 1966;5:264–76. [PubMed] [Google Scholar]
  • 48.Braun RJ. Dynamics of the tear film. Annu Rev Fluid Mech. 2012;44:267, 97. [Google Scholar]
  • 49.Dartt DA. Regulation of mucin and fluid secretion by conjunctival epithelial cells. Prog Retin Eye Res. 2002;21:555–76. doi: 10.1016/s1350-9462(02)00038-1. [DOI] [PubMed] [Google Scholar]
  • 50.Maitchouk DY, Beuerman RW, Ohta T, Stern M, Varnell RJ. Tear production after unilateral removal of the main lacrimal gland in squirrel monkeys. Arch Ophthalmol. 2000;118:246–52. doi: 10.1001/archopht.118.2.246. [DOI] [PubMed] [Google Scholar]
  • 51.Botelho SY, Hisada M, Fuenmayor N. Functional innervation of the lacrimal gland in the cat. Origin of secretomotor fibers in the lacrimal nerve. Arch Ophthalmol. 1966;76:581–8. doi: 10.1001/archopht.1966.03850010583019. [DOI] [PubMed] [Google Scholar]
  • 52.Sibony PA, Walcott B, McKeon C, Jakobiec FA. Vasoactive intestinal poly-peptide and the innervation of the human lacrimal gland. Arch Ophthalmol. 1988;106:1085–8. doi: 10.1001/archopht.1988.01060140241033. [DOI] [PubMed] [Google Scholar]
  • 53.Botelho SY. Tears and the lacrimal gland. Sci Am. 1964;211:78–86. doi: 10.1038/scientificamerican1064-78. [DOI] [PubMed] [Google Scholar]
  • 54.Craig JP, Willcox MD, Argueso P, Maissa C, Stahl U, Tomlinson A, et al. The TFOS International Workshop on Contact Lens Discomfort: report of the contact lens interactions with the tear film subcommittee. Invest Ophthalmol Vis Sci. 2013;54:TFOS123–56. doi: 10.1167/iovs.13-13235. [DOI] [PubMed] [Google Scholar]
  • 55.Belmonte C, Nichols JJ, Cox SM, Brock JA, Begley CG, Bereiter DA, et al. TFOS DEWS II Pain and Sensation report. Ocul Surf. 2017;15:404–37. doi: 10.1016/j.jtos.2017.05.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Brown SI, Dervichian DG. Hydrodynamics of blinking. In vitro study of the interaction of the superficial oily layer and the tears. Arch Ophthalmol. 1969;82:541–7. doi: 10.1001/archopht.1969.00990020543020. [DOI] [PubMed] [Google Scholar]
  • 57.King-Smith PE, Fink BA, Nichols JJ, Nichols KK, Braun RJ, McFadden GB. The contribution of lipid layer movement to tear film thinning and breakup. Invest Ophthalmol Vis Sci. 2009;50:2747–56. doi: 10.1167/iovs.08-2459. [DOI] [PubMed] [Google Scholar]
  • 58.Bron AJ, Yokoi N, Gaffney EA, Tiffany JM. A solute gradient in the tear meniscus. I. A hypothesis to explain Marx’s line. Ocul Surf. 2011;9:70–91. doi: 10.1016/s1542-0124(11)70014-3. [DOI] [PubMed] [Google Scholar]
  • 59.Bron AJ, Yokoi N, Gaffney EA, Tiffany JM. A solute gradient in the tear meniscus. II. Implications for lid margin disease, including meibomian gland dysfunction. Ocul Surf. 2011;9:92–7. doi: 10.1016/s1542-0124(11)70015-5. [DOI] [PubMed] [Google Scholar]
  • 60.Maurice DM. The dynamics and drainage of tears. Int Ophthalmol Clin. 1973;13:103–16. doi: 10.1097/00004397-197301310-00009. [DOI] [PubMed] [Google Scholar]
  • 61.Khanal S, Millar TJ. Nanoscale phase dynamics of the normal tear film. Nanomedicine. 2010;6:707–13. doi: 10.1016/j.nano.2010.06.002. [DOI] [PubMed] [Google Scholar]
  • 62.Kuppens EV, Stolwijk TR, de Keizer RJ, van Best JA. Basal tear turnover and topical timolol in glaucoma patients and healthy controls by fluorophotometry. Invest Ophthalmol Vis Sci. 1992;33:3442–8. [PubMed] [Google Scholar]
  • 63.Kuppens EV, van Best JA, Sterk CC, de Keizer RJ. Decreased basal tear turnover in patients with untreated primary open-angle glaucoma. Am J Ophthalmol. 1995;120:41–6. doi: 10.1016/s0002-9394(14)73757-2. [DOI] [PubMed] [Google Scholar]
  • 64.van Best JA, Benitez del Castillo JM, Coulangeon LM. Measurement of basal tear turnover using a standardized protocol. European concerted action on ocular fluorometry. Graefes Arch Clin Exp Ophthalmol. 1995;233:1–7. doi: 10.1007/BF00177778. [DOI] [PubMed] [Google Scholar]
  • 65.Nichols JJ, Mitchell GL, King-Smith PE. Thinning rate of the precorneal and prelens tear films. Invest Ophthalmol Vis Sci. 2005;46:2353–61. doi: 10.1167/iovs.05-0094. [DOI] [PubMed] [Google Scholar]
  • 66.Rolland JP, Huang J, Hindman HB. In vivo measurement of tear film dynamics with optical coherence tomography and maximum-likelihood estimation. Invest Ophthalmol Vis Sci. 2016;57 E-Abstract 6179. [Google Scholar]
  • 67.Craig JP, Tomlinson A. Importance of the lipid layer in human tear film stability and evaporation. Optom Vis Sci. 1997;74:8–13. doi: 10.1097/00006324-199701000-00014. [DOI] [PubMed] [Google Scholar]
  • 68.Gaffney EA, Tiffany JM, Yokoi N, Bron AJ. A mass and solute balance model for tear volume and osmolarity in the normal and the dry eye. Prog Retin Eye Res. 2010;29:59–78. doi: 10.1016/j.preteyeres.2009.11.002. [DOI] [PubMed] [Google Scholar]
  • 69.Sullivan BD, Crews LA, Messmer EM, Foulks GN, Nichols KK, Baenninger P, et al. Correlations between commonly used objective signs and symptoms for the diagnosis of dry eye disease: clinical implications. Acta Ophthalmol. 2014;92:161–6. doi: 10.1111/aos.12012. [DOI] [PubMed] [Google Scholar]
  • 70.de Monchy I, Gendron G, Miceli C, Pogorzalek N, Mariette X, Labetoulle M. Combination of the Schirmer I and phenol red thread tests as a rescue strategy for diagnosis of ocular dryness associated with Sjogren’s syndrome. Invest Ophthalmol Vis Sci. 2011;52:5167–73. doi: 10.1167/iovs.10-6671. [DOI] [PubMed] [Google Scholar]
  • 71.Hamano T, Mitsunaga S, Kotani S, Hamano T, Hamano K, Hamano H, et al. Tear volume in relation to contact lens wear and age. CLAO J. 1990;16:57–61. [PubMed] [Google Scholar]
  • 72.Wright JC, Meger GE. A review of the Schirmer test for tear production. Arch Ophthalmol. 1962;67:564–5. doi: 10.1001/archopht.1962.00960020564008. [DOI] [PubMed] [Google Scholar]
  • 73.Mainstone JC, Bruce AS, Golding TR. Tear meniscus measurement in the diagnosis of dry eye. Curr Eye Res. 1996;15:653–61. doi: 10.3109/02713689609008906. [DOI] [PubMed] [Google Scholar]
  • 74.Yokoi N, Bron AJ, Tiffany JM, Kinoshita S. Reflective meniscometry: a new field of dry eye assessment. Cornea. 2000;19:S37–43. doi: 10.1097/00003226-200005001-00008. [DOI] [PubMed] [Google Scholar]
  • 75.Xu KP, Tsubota K. Correlation of tear clearance rate and fluorophotometric assessment of tear turnover. Br J Ophthalmol. 1995;79:1042–5. doi: 10.1136/bjo.79.11.1042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Xu KP, Yagi Y, Toda I, Tsubota K. Tear function index. A new measure of dry eye. Arch Ophthalmol. 1995;113:84–8. doi: 10.1001/archopht.1995.01100010086025. [DOI] [PubMed] [Google Scholar]
  • 77.Korb DR, Greiner JV, Glonek T, Whalen A, Hearn SL, Esway JE, et al. Human and rabbit lipid layer and interference pattern observations. Adv Exp Med Biol. 1998;438:305–8. doi: 10.1007/978-1-4615-5359-5_42. [DOI] [PubMed] [Google Scholar]
  • 78.King-Smith PE, Nichols JJ, Nichols KK, Fink BA, Braun RJ. Contributions of evaporation and other mechanisms to tear film thinning and break-up. Optom Vis Sci. 2008;85:623–30. doi: 10.1097/OPX.0b013e318181ae60. [DOI] [PubMed] [Google Scholar]
  • 79.King-Smith PE, Ramamoorthy P, Braun RJ, Nichols JJ. Tear film images and breakup analyzed using fluorescent quenching. Invest Ophthalmol Vis Sci. 2013;54:6003–11. doi: 10.1167/iovs.13-12628. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Papas E. Tear break-up time: clinical procedures and their effects. Ophthalmic Physiol Opt. 1999;19:274–5. doi: 10.1046/j.1475-1313.1999.00422.x. [DOI] [PubMed] [Google Scholar]
  • 81.Cox SM, Nichols KK, Nichols JJ. Agreement between automated and traditional measures of tear film breakup. Optom Vis Sci. 2015;92:e257–63. doi: 10.1097/OPX.0000000000000648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Nichols KK, Mitchell GL, Zadnik K. The repeatability of clinical measurements of dry eye. Cornea. 2004;23:273–85. doi: 10.1097/00003226-200404000-00010. [DOI] [PubMed] [Google Scholar]
  • 83.Cho P, Brown B, Chan I, Conway R, Yap M. Reliability of the tear break-up time technique of assessing tear stability and the locations of the tear breakup in Hong Kong Chinese. Optom Vis Sci. 1992;69:879–85. doi: 10.1097/00006324-199211000-00007. [DOI] [PubMed] [Google Scholar]
  • 84.Abelson R, Lane KJ, Rodriguez J, Johnston P, Angjeli E, Ousler G, et al. A single-center study evaluating the effect of the controlled adverse environment (CAE(SM)) model on tear film stability. Clin Ophthalmol. 2012;6:1865–72. doi: 10.2147/OPTH.S33905. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Pult H, Riede-Pult BH. A new modified fluorescein strip: its repeatability and usefulness in tear film break-up time analysis. Cont Lens Anterior Eye. 2012;35:35–8. doi: 10.1016/j.clae.2011.07.005. [DOI] [PubMed] [Google Scholar]
  • 86.Korb DR, Greiner JV, Herman J. Comparison of fluorescein break-up time measurement reproducibility using standard fluorescein strips versus the Dry Eye Test (DET) method. Cornea. 2001;20:811–5. doi: 10.1097/00003226-200111000-00007. [DOI] [PubMed] [Google Scholar]
  • 87.Lamble JW, Gilbert D, Ashford JJ. The break-up time of artificial pre-ocular films on the rabbit cornea. J Pharm Pharmacol. 1976;28:450–1. doi: 10.1111/j.2042-7158.1976.tb04654.x. [DOI] [PubMed] [Google Scholar]
  • 88.Mengher LS, Bron AJ, Tonge SR, Gilbert DJ. A non-invasive instrument for clinical assessment of the pre-corneal tear film stability. Curr Eye Res. 1985;4:1–7. doi: 10.3109/02713688508999960. [DOI] [PubMed] [Google Scholar]
  • 89.Patel S, Murray D, McKenzie A, Shearer DS, McGrath BD. Effects of fluorescein on tear breakup time and on tear thinning time. Am J Optom Physiol Opt. 1985;62:188–90. doi: 10.1097/00006324-198503000-00006. [DOI] [PubMed] [Google Scholar]
  • 90.Hirji N, Patel S, Callander M. Human tear film pre-rupture phase time (TP-RPT)–a non-invasive technique for evaluating the pre-corneal tear film using a novel keratometer mire. Ophthalmic Physiol Opt. 1989;9:139–42. doi: 10.1111/j.1475-1313.1989.tb00833.x. [DOI] [PubMed] [Google Scholar]
  • 91.Alonso-Caneiro D, Iskander DR, Collins MJ. Tear film surface quality with soft contact lenses using dynamic-area high-speed videokeratoscopy. Eye Contact Lens. 2009;35:227–31. doi: 10.1097/ICL.0b013e3181b3350f. [DOI] [PubMed] [Google Scholar]
  • 92.Downie LE. Automated tear film surface quality breakup time as a novel clinical marker for tear hyperosmolarity in dry eye disease. Invest Ophthalmol Vis Sci. 2015;56:7260–8. doi: 10.1167/iovs.15-17772. [DOI] [PubMed] [Google Scholar]
  • 93.Hirota M, Uozato H, Kawamorita T, Shibata Y, Yamamoto S. Effect of incomplete blinking on tear film stability. Optom Vis Sci. 2013;90:650–7. doi: 10.1097/OPX.0b013e31829962ec. [DOI] [PubMed] [Google Scholar]
  • 94.Hong J, Sun X, Wei A, Cui X, Li Y, Qian T, et al. Assessment of tear film stability in dry eye with a newly developed keratograph. Cornea. 2013;32:716–21. doi: 10.1097/ICO.0b013e3182714425. [DOI] [PubMed] [Google Scholar]
  • 95.Koh S, Ikeda C, Fujimoto H, Oie Y, Soma T, Maeda N, et al. Regional differences in tear film stability and meibomian glands in patients with aqueous-deficient dry eye. Eye Contact Lens. 2016;42:250–5. doi: 10.1097/ICL.0000000000000191. [DOI] [PubMed] [Google Scholar]
  • 96.Kopf M, Yi F, Robert Iskander D, Collins MJ, Shaw AJ, Straker B. Tear film surface quality with soft contact lenses using dynamic videokeratoscopy. J Opt. 2008;1:14–21. [Google Scholar]
  • 97.Gumus K, Crockett CH, Rao K, Yeu E, Weikert MP, Shirayama M, et al. Noninvasive assessment of tear stability with the tear stability analysis system in tear dysfunction patients. Invest Ophthalmol Vis Sci. 2011;52:456–61. doi: 10.1167/iovs.10-5292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Goto T, Zheng X, Okamoto S, Ohashi Y. Tear film stability analysis system: introducing a new application for videokeratography. Cornea. 2004;23:S65–70. doi: 10.1097/01.ico.0000136685.88489.70. [DOI] [PubMed] [Google Scholar]
  • 99.Wolffsohn JS, Arita R, Chalmers R, Djalilian A, Dogru M, Dumbleton K, et al. TFOS DEWS II Diagnostic Methodology report. Ocul Surf. 2017;15:539–74. doi: 10.1016/j.jtos.2017.05.001. [DOI] [PubMed] [Google Scholar]
  • 100.Su TY, Ho WT, Lu CY, Chang SW, Chiang HK. Correlations among ocular surface temperature difference value, the tear meniscus height, Schirmer’s test and fluorescein tear film break up time. Br J Ophthalmol. 2015;99:482–7. doi: 10.1136/bjophthalmol-2014-305183. [DOI] [PubMed] [Google Scholar]
  • 101.Szczesna DH, Jaronski J, Kasprzak HT, Stenevi U. Interferometric measurements of dynamic changes of tear film. J Biomed Opt. 2006;11:34028. doi: 10.1117/1.2209881. [DOI] [PubMed] [Google Scholar]
  • 102.Szczesna DH, Kasprzak HT, Jaronski J, Rydz A, Stenevi U. An interferometric method for the dynamic evaluation of the tear film. Acta Ophthalmol Scand. 2007;85:202–8. doi: 10.1111/j.1600-0420.2006.00802.x. [DOI] [PubMed] [Google Scholar]
  • 103.Guillon JP. Tear film structure and contact lenses. In: Holly FJ, editor. The pre-ocular tear film in health, disease and contact lens wear. Lubbock, TX: Dry Eye Institute; 1986. pp. 914–35. [Google Scholar]
  • 104.Doane MG. An instrument for in vivo tear film interferometry. Optom Vis Sci. 1989;66:383–8. doi: 10.1097/00006324-198906000-00008. [DOI] [PubMed] [Google Scholar]
  • 105.Szczesna DH, Iskander DR. Robust estimation of tear film surface quality in lateral shearing interferometry. J Biomed Opt. 2009;14:064039. doi: 10.1117/1.3275474. [DOI] [PubMed] [Google Scholar]
  • 106.Szczesna DH, Alonso-Caneiro D, Iskander DR, Read SA, Collins MJ. Predicting dry eye using noninvasive techniques of tear film surface assessment. Invest Ophthalmol Vis Sci. 2011;52:751–6. doi: 10.1167/iovs.10-5173. [DOI] [PubMed] [Google Scholar]
  • 107.Benito A, Perez GM, Mirabet S, Vilaseca M, Pujol J, Marin JM, et al. Objective optical assessment of tear-film quality dynamics in normal and mildly symptomatic dry eyes. J Cataract Refract Surg. 2011;37:1481–7. doi: 10.1016/j.jcrs.2011.03.036. [DOI] [PubMed] [Google Scholar]
  • 108.Kobashi H, Kamiya K, Yanome K, Igarashi A, Shimizu K. Longitudinal assessment of optical quality and intraocular scattering using the double-pass instrument in normal eyes and eyes with short tear breakup time. PLoS One. 2013;8:e82427. doi: 10.1371/journal.pone.0082427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Su TY, Ho WT, Chang SW, Chiang HK. Thermographic evaluation of tear film break-up time to study tear film stability. Int J Therm Sci. 2016;99:36–40. [Google Scholar]
  • 110.Rieger G. The importance of the precorneal tear film for the quality of optical imaging. Br J Ophthalmol. 1992;76:157–8. doi: 10.1136/bjo.76.3.157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Goto E, Yagi Y, Matsumoto Y, Tsubota K. Impaired functional visual acuity of dry eye patients. Am J Ophthalmol. 2002;133:181–6. doi: 10.1016/s0002-9394(01)01365-4. [DOI] [PubMed] [Google Scholar]
  • 112.Denoyer A, Rabut G, Baudouin C. Tear film aberration dynamics and vision-related quality of life in patients with dry eye disease. Ophthalmology. 2012;119:1811–8. doi: 10.1016/j.ophtha.2012.03.004. [DOI] [PubMed] [Google Scholar]
  • 113.Liu H, Thibos L, Begley CG, Bradley A. Measurement of the time course of optical quality and visual deterioration during tear break-up. Invest Ophthalmol Vis Sci. 2010;51:3318–26. doi: 10.1167/iovs.09-4831. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Tomlinson A, Khanal S. Assessment of tear film dynamics: quantification approach. Ocul Surf. 2005;3:81–95. doi: 10.1016/s1542-0124(12)70157-x. [DOI] [PubMed] [Google Scholar]
  • 115.Lemp MA. Report of the national eye institute/industry workshop on clinical trials in dry eyes. CLAO J. 1995;21:221–32. [PubMed] [Google Scholar]
  • 116.The definition and classification of dry eye disease: report of the Definition and Classification Subcommittee of the International Dry Eye Workshop (2007) Ocul Surf. 2007;5:75–92. doi: 10.1016/s1542-0124(12)70081-2. [DOI] [PubMed] [Google Scholar]
  • 117.Murube J. Tear osmolarity. Ocul Surf. 2006;4:62–73. doi: 10.1016/s1542-0124(12)70028-9. [DOI] [PubMed] [Google Scholar]
  • 118.Versura P, Campos EC. TearLab(R) Osmolarity System for diagnosing dry eye. Expert Rev Mol Diagn. 2013;13:119–29. doi: 10.1586/erm.12.142. [DOI] [PubMed] [Google Scholar]
  • 119.Stahl U, Willcox M, Stapleton F. Osmolality and tear film dynamics. Clin Exp Optom. 2012;95:3–11. doi: 10.1111/j.1444-0938.2011.00634.x. [DOI] [PubMed] [Google Scholar]
  • 120.Tomlinson A, Bron AJ, Korb DR, Amano S, Paugh JR, Pearce EI, et al. The international workshop on meibomian gland dysfunction: report of the diagnosis subcommittee. Invest Ophthalmol Vis Sci. 2011;52:2006–49. doi: 10.1167/iovs.10-6997f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Srinivasan S, Nichols KK. Collecting tear osmolarity measurements in the diagnosis of dry eye. Expert Rev Ophthalmol. 2009;4:451–3. [Google Scholar]
  • 122.Messmer EM, Bulgen M, Kampik A. Hyperosmolarity of the tear film in dry eye syndrome. Dev Ophthalmol. 2010;45:129–38. doi: 10.1159/000315026. [DOI] [PubMed] [Google Scholar]
  • 123.Khanal S, Tomlinson A, Diaper CJ. Tear physiology of aqueous deficiency and evaporative dry eye. Optom Vis Sci. 2009;86:1235–40. doi: 10.1097/OPX.0b013e3181bc63cc. [DOI] [PubMed] [Google Scholar]
  • 124.Srinivasan S, Joyce E, Jones LW. Tear osmolality and ferning patterns in postmenopausal women. Optom Vis Sci. 2007;84:588–92. doi: 10.1097/OPX.0b013e3180dc9a23. [DOI] [PubMed] [Google Scholar]
  • 125.Tomlinson A, McCann LC, Pearce EI. Comparison of human tear film osmolarity measured by electrical impedance and freezing point depression techniques. Cornea. 2010;29:1036–41. doi: 10.1097/ICO.0b013e3181cd9a1d. [DOI] [PubMed] [Google Scholar]
  • 126.Versura P, Profazio V, Campos EC. Performance of tear osmolarity compared to previous diagnostic tests for dry eye diseases. Curr Eye Res. 2010;35:553–64. doi: 10.3109/02713683.2010.484557. [DOI] [PubMed] [Google Scholar]
  • 127.Jacobi C, Jacobi A, Kruse FE, Cursiefen C. Tear film osmolarity measurements in dry eye disease using electrical impedance technology. Cornea. 2011;30:1289–92. doi: 10.1097/ICO.0b013e31821de383. [DOI] [PubMed] [Google Scholar]
  • 128.Lemp MA, Bron AJ, Baudouin C, Benitez Del Castillo JM, Geffen D, Tauber J, et al. Tear osmolarity in the diagnosis and management of dry eye disease. Am J Ophthalmol. 2011;151:792–8. doi: 10.1016/j.ajo.2010.10.032. [DOI] [PubMed] [Google Scholar]
  • 129.Szalai E, Berta A, Szekanecz Z, Szucs G, Modis L., Jr Evaluation of tear osmolarity in non-Sjogren and Sjogren syndrome dry eye patients with the TearLab system. Cornea. 2012;31:867–71. doi: 10.1097/ICO.0b013e3182532047. [DOI] [PubMed] [Google Scholar]
  • 130.Li M, Du C, Zhu D, Shen M, Cui L, Wang J. Daytime variations of tear osmolarity and tear meniscus volume. Eye Contact Lens. 2012;38:282–7. doi: 10.1097/ICL.0b013e31825fed57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Chen SP, Massaro-Giordano G, Pistilli M, Schreiber CA, Bunya VY. Tear osmolarity and dry eye symptoms in women using oral contraception and contact lenses. Cornea. 2013;32:423–8. doi: 10.1097/ICO.0b013e3182662390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Hassan Z, Szalai E, Berta A, Modis L, Jr, Nemeth G. Assessment of tear osmolarity and other dry eye parameters in post-LASIK eyes. Cornea. 2013;32:e142–5. doi: 10.1097/ICO.0b013e318290496d. [DOI] [PubMed] [Google Scholar]
  • 133.Keech A, Senchyna M, Jones L. Impact of time between collection and collection method on human tear fluid osmolarity. Curr Eye Res. 2013;38:428–36. doi: 10.3109/02713683.2013.763987. [DOI] [PubMed] [Google Scholar]
  • 134.Garcia N, Teson M, Enriquez-de-Salamanca A, Mena L, Sacristan A, Fernandez I, et al. Basal values, intra-day and inter-day variations in tear film osmolarity and tear fluorescein clearance. Curr Eye Res. 2014;39:673–9. doi: 10.3109/02713683.2013.865757. [DOI] [PubMed] [Google Scholar]
  • 135.Koktekir BE, Bozkurt B, Gonul S, Gedik S, Okudan S. Effect of religious fasting on tear osmolarity and ocular surface. Eye Contact Lens. 2014;40:239–42. doi: 10.1097/ICL.0000000000000044. [DOI] [PubMed] [Google Scholar]
  • 136.Masmali A, Alrabiah S, Alharbi A, El-Hiti GA, Almubrad T. Investigation of tear osmolarity using the TearLab osmolarity system in normal adults in Saudi Arabia. Eye Contact Lens. 2014;40:74–8. doi: 10.1097/ICL.0000000000000002. [DOI] [PubMed] [Google Scholar]
  • 137.Bunya VY, Fuerst NM, Pistilli M, McCabe BE, Salvo R, Macchi I, et al. Variability of tear osmolarity in patients with dry eye. JAMA Ophthalmol. 2015;133:662–7. doi: 10.1001/jamaophthalmol.2015.0429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Tomlinson A, Khanal S, Ramaesh K, Diaper C, McFadyen A. Tear film osmolarity: determination of a referent for dry eye diagnosis. Invest Ophthalmol Vis Sci. 2006;47:4309–15. doi: 10.1167/iovs.05-1504. [DOI] [PubMed] [Google Scholar]
  • 139.Caffery B, Chalmers RL, Marsden H, Nixon G, Watanabe R, Harrison W, et al. Correlation of tear osmolarity and dry eye symptoms in convention attendees. Optom Vis Sci. 2014;91:142–9. doi: 10.1097/OPX.0000000000000130. [DOI] [PubMed] [Google Scholar]
  • 140.Fuerst N, Langelier N, Massaro-Giordano M, Pistilli M, Stasi K, Burns C, et al. Tear osmolarity and dry eye symptoms in diabetics. Clin Ophthalmol. 2014;8:507–15. doi: 10.2147/OPTH.S51514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Eperjesi F, Aujla M, Bartlett H. Reproducibility and repeatability of the OcuSense TearLab osmometer. Graefes Arch Clin Exp Ophthalmol. 2012;250:1201–5. doi: 10.1007/s00417-012-1961-4. [DOI] [PubMed] [Google Scholar]
  • 142.Niimi J, Tan B, Chang J, Zhou Y, Ghanekar A, Wong M, et al. Diurnal pattern of tear osmolarity and its relationship to corneal thickness and deswelling. Cornea. 2013;32:1305–10. doi: 10.1097/ICO.0b013e31829b21d1. [DOI] [PubMed] [Google Scholar]
  • 143.Yeh TN, Graham AD, Lin MC. Relationships among tear film stability, osmolarity, and dryness symptoms. Optom Vis Sci. 2015;92:e264–72. doi: 10.1097/OPX.0000000000000649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Tomlinson A, Pearce EI, Simmons PA, Blades K. Effect of oral contraceptives on tear physiology. Ophthalmic Physiol Opt. 2001;21:9–16. [PubMed] [Google Scholar]
  • 145.Vehof J, Wang B, Kozareva D, Hysi PG, Snieder H, Hammond CJ. The heritability of dry eye disease in a female twin cohort. Invest Ophthalmol Vis Sci. 2014;55:7278–83. doi: 10.1167/iovs.14-15200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Sullivan BD, Pepose JS, Foulks GN. Progressively increased variation in tear osmolarity mirrors dry eye severity. JAMA Ophthalmol. 2015;133:1481–2. doi: 10.1001/jamaophthalmol.2015.3854. [DOI] [PubMed] [Google Scholar]
  • 147.Lee YB, Koh JW, Hyon JY, Wee WR, Kim JJ, Shin YJ. Sleep deprivation reduces tear secretion and impairs the tear film. Invest Ophthalmol Vis Sci. 2014;55:3525–31. doi: 10.1167/iovs.14-13881. [DOI] [PubMed] [Google Scholar]
  • 148.Willmann G, Schatz A, Fischer MD, Schommer K, Zrenner E, Bartz-Schmidt KU, et al. Exposure to high altitude alters tear film osmolarity and breakup time. High Alt Med Biol. 2014;15:203–7. doi: 10.1089/ham.2013.1103. [DOI] [PubMed] [Google Scholar]
  • 149.Khanal S, Millar TJ. Barriers to clinical uptake of tear osmolarity measurements. Br J Ophthalmol. 2012;96:341–4. doi: 10.1136/bjo.2011.202754. [DOI] [PubMed] [Google Scholar]
  • 150.Fortes MB, Diment BC, Di Felice U, Gunn AE, Kendall JL, Esmaeelpour M, et al. Tear fluid osmolarity as a potential marker of hydration status. Med Sci Sports Exerc. 2011;43:1590–7. doi: 10.1249/MSS.0b013e31820e7cb6. [DOI] [PubMed] [Google Scholar]
  • 151.Walsh NP, Fortes MB, Esmaeelpour M. Influence of modest changes in whole-body hydration on tear fluid osmolarity: important considerations for dry eye disease detection. Cornea. 2011;30:1517. doi: 10.1097/ICO.0b013e31821ddd3a. author reply -8. [DOI] [PubMed] [Google Scholar]
  • 152.Walsh NP, Fortes MB, Raymond-Barker P, Bishop C, Owen J, Tye E, et al. Is whole-body hydration an important consideration in dry eye? Invest Ophthalmol Vis Sci. 2012;53:6622–7. doi: 10.1167/iovs.12-10175. [DOI] [PubMed] [Google Scholar]
  • 153.Braun RJ, King-Smith PE, Begley CG, Li L, Gewecke NR. Dynamics and function of the tear film in relation to the blink cycle. Prog Retin Eye Res. 2015;45:132–64. doi: 10.1016/j.preteyeres.2014.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Li L, Braun RJ, Driscoll TA, Henshaw WD, Banks JW, King-Smith PE. Computed tear film and osmolarity dynamics on an eye-shaped domain. Math Med Biol. 2016;33:123–57. doi: 10.1093/imammb/dqv013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Begley C, Simpson T, Liu H, Salvo E, Wu Z, Bradley A, et al. Quantitative analysis of tear film fluorescence and discomfort during tear film instability and thinning. Invest Ophthalmol Vis Sci. 2013;54:2645–53. doi: 10.1167/iovs.12-11299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Varikooty J, Simpson TL. The interblink interval I: the relationship between sensation intensity and tear film disruption. Invest Ophthalmol Vis Sci. 2009;50:1087–92. doi: 10.1167/iovs.08-1843. [DOI] [PubMed] [Google Scholar]
  • 157.Liu H, Begley C, Chen M, Bradley A, Bonanno J, McNamara NA, et al. A link between tear instability and hyperosmolarity in dry eye. Invest Ophthalmol Vis Sci. 2009;50:3671–9. doi: 10.1167/iovs.08-2689. [DOI] [PubMed] [Google Scholar]
  • 158.Craig JP, Singh I, Tomlinson A, Morgan PB, Efron N. The role of tear physiology in ocular surface temperature. Eye (Lond) 2000;14(Pt 4):635–41. doi: 10.1038/eye.2000.156. [DOI] [PubMed] [Google Scholar]
  • 159.Research in dry eye: report of the research subcommittee of the international dry eye WorkShop (2007) Ocul Surf. 2007;5:179–93. doi: 10.1016/s1542-0124(12)70086-1. [DOI] [PubMed] [Google Scholar]
  • 160.Suzuki M, Massingale ML, Ye F, Godbold J, Elfassy T, Vallabhajosyula M, et al. Tear osmolarity as a biomarker for dry eye disease severity. Invest Ophthalmol Vis Sci. 2010;51:4557–61. doi: 10.1167/iovs.09-4596. [DOI] [PubMed] [Google Scholar]
  • 161.Sullivan BD, Crews LA, Sonmez B, de la Paz MF, Comert E, Charoenrook V, et al. Clinical utility of objective tests for dry eye disease: variability over time and implications for clinical trials and disease management. Cornea. 2012;31:1000–8. doi: 10.1097/ICO.0b013e318242fd60. [DOI] [PubMed] [Google Scholar]
  • 162.Sullivan BD, Whitmer D, Nichols KK, Tomlinson A, Foulks GN, Geerling G, et al. An objective approach to dry eye disease severity. Invest Ophthalmol Vis Sci. 2010;51:6125–30. doi: 10.1167/iovs.10-5390. [DOI] [PubMed] [Google Scholar]
  • 163.Tukenmez-Dikmen N, Yildiz EH, Imamoglu S, Turan-Vural E, Sevim MS. Correlation of dry eye workshop dry eye severity grading system with tear meniscus measurement by optical coherence tomography and tear osmolarity. Eye Contact Lens. 2016;42:153–7. doi: 10.1097/ICL.0000000000000168. [DOI] [PubMed] [Google Scholar]
  • 164.Utine CA, Bicakcigil M, Yavuz S, Ciftci F. Tear osmolarity measurements in dry eye related to primary Sjogren’s syndrome. Curr Eye Res. 2011;36:683–90. doi: 10.3109/02713683.2011.571357. [DOI] [PubMed] [Google Scholar]
  • 165.Gilbard JP, Farris RL, Santamaria J., 2nd Osmolarity of tear microvolumes in keratoconjunctivitis sicca. Arch Ophthalmol. 1978;96:677–81. doi: 10.1001/archopht.1978.03910050373015. [DOI] [PubMed] [Google Scholar]
  • 166.Alves M, Reinach PS, Paula JS, Vellasco e Cruz AA, Bachette L, Faustino J, et al. Comparison of diagnostic tests in distinct well-defined conditions related to dry eye disease. PLoS One. 2014;9:e97921. doi: 10.1371/journal.pone.0097921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Bunya VY, Langelier N, Chen S, Pistilli M, Vivino FB, Massaro-Giordano G. Tear osmolarity in Sjogren syndrome. Cornea. 2013;32:922–7. doi: 10.1097/ICO.0b013e31827e2a5e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Amparo F, Jin Y, Hamrah P, Schaumberg DA, Dana R. What is the value of incorporating tear osmolarity measurement in assessing patient response to therapy in dry eye disease? Am J Ophthalmol. 2014;157:69–77. doi: 10.1016/j.ajo.2013.07.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Berchicci L, Iuliano L, Miserocchi E, Bandello F, Modorati G. Tear osmolarity in ocular graft-versus-host disease. Cornea. 2014;33:1252–6. doi: 10.1097/ICO.0000000000000283. [DOI] [PubMed] [Google Scholar]
  • 170.Na KS, Yoo YS, Hwang KY, Mok JW, Joo CK. Tear osmolarity and ocular surface parameters as diagnostic markers of ocular graft-versus-host disease. Am J Ophthalmol. 2015;160:143–9.e1. doi: 10.1016/j.ajo.2015.04.002. [DOI] [PubMed] [Google Scholar]
  • 171.Schargus M, Meyer-ter-Vehn T, Menrath J, Grigoleit GU, Geerling G. Correlation between tear film osmolarity and the disease score of the International Chronic Ocular Graft-Versus-Host-Disease Consensus Group in hematopoietic stem cell transplantation patients. Cornea. 2015;34:911–6. doi: 10.1097/ICO.0000000000000494. [DOI] [PubMed] [Google Scholar]
  • 172.Rolando M, Baldi F, Calabria GA. Tear mucus ferning in keratoconjunctivitis sicca. In: Holly FJ, editor. The preocular tear film in health, disease and contact lens wear. Lubbock, Tex: Dry Eye Institute; 1986. pp. 203–9. [Google Scholar]
  • 173.Kogbe O, Liotet S, Tiffany JM. Factors responsible for tear ferning. Cornea. 1991;10:433–44. doi: 10.1097/00003226-199109000-00013. [DOI] [PubMed] [Google Scholar]
  • 174.Golding TR, Brennan NA. The basis of tear ferning. Clin Exp Optom. 1989;72:102–12. [Google Scholar]
  • 175.Norn M. Quantitative tear ferning. Clinical investigations. Acta Ophthalmol (Copenh) 1994;72:369–72. doi: 10.1111/j.1755-3768.1994.tb02775.x. [DOI] [PubMed] [Google Scholar]
  • 176.Vaikoussis E, Georgiou P, Nomicarios D. Tear mucus ferning in patients with Sjogren’s syndrome. Doc Ophthalmol. 1994;87:145–51. doi: 10.1007/BF01204791. [DOI] [PubMed] [Google Scholar]
  • 177.Maragou M, Vaikousis E, Ntre A, Koronis N, Georgiou P, Hatzidimitriou E, et al. Tear and saliva ferning tests in Sjogren’s syndrome (SS) Clin Rheumatol. 1996;15:125–32. doi: 10.1007/BF02230328. [DOI] [PubMed] [Google Scholar]
  • 178.Masmali AM, Al-Qhtani S, Al-Gasham TM, El-Hiti GA, Purslow C, Murphy PJ. Application of a new grading scale for tear ferning in non-dry eye and dry eye subjects. Cont Lens Anterior Eye. 2015;38:39–43. doi: 10.1016/j.clae.2014.09.007. [DOI] [PubMed] [Google Scholar]
  • 179.Bitton E, Keech A, Jones L, Simpson T. Subjective and objective variation of the tear film pre- and post-sleep. Optom Vis Sci. 2008;85:740–9. doi: 10.1097/OPX.0b013e318181a92f. [DOI] [PubMed] [Google Scholar]
  • 180.Masmali AM, Al-Bahlal JM, El-Hiti GA, Akhtar S, Purslow C, Murphy PJ, et al. Repeatability and diurnal variation of tear ferning test. Eye Contact Lens. 2015;41:262–7. doi: 10.1097/ICL.0000000000000116. [DOI] [PubMed] [Google Scholar]
  • 181.Sharanjeet K, Ho CY, Mutalib HA, Ghazali AR. The relationship between tear ferning patterns and non-invasive tear break-up time in normal Asian population. J Optom. 2016;9:175–81. doi: 10.1016/j.optom.2015.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Versura P, Fresina M, Campos EC. Ocular surface changes over the menstrual cycle in women with and without dry eye. Gynecol Endocrinol. 2007;23:385–90. doi: 10.1080/09513590701350390. [DOI] [PubMed] [Google Scholar]
  • 183.Cennamo GL, Del Prete A, Forte R, Cafiero G, Del Prete S, Marasco D. Impression cytology with scanning electron microscopy: a new method in the study of conjunctival microvilli. Eye (Lond) 2008;22:138–43. doi: 10.1038/sj.eye.6702873. [DOI] [PubMed] [Google Scholar]
  • 184.Ravazzoni L, Ghini C, Macri A, Rolando M. Forecasting of hydrophilic contact lens tolerance by means of tear ferning test. Graefes Arch Clin Exp Ophthalmol. 1998;236:354–8. doi: 10.1007/s004170050090. [DOI] [PubMed] [Google Scholar]
  • 185.Pensyl CD, Dillehay SM. The repeatability of tear mucus ferning grading. Optom Vis Sci. 1998;75:600–4. doi: 10.1097/00006324-199808000-00025. [DOI] [PubMed] [Google Scholar]
  • 186.Evans KS, North RV, Purslow C. Tear ferning in contact lens wearers. Ophthalmic Physiol Opt. 2009;29:199–204. doi: 10.1111/j.1475-1313.2008.00626.x. [DOI] [PubMed] [Google Scholar]
  • 187.Julio G, Lluch S, Cardona G, Fornieles A, Merindano D. Item by item analysis strategy of the relationship between symptoms and signs in early dry eye. Curr Eye Res. 2012;37:357–64. doi: 10.3109/02713683.2012.654884. [DOI] [PubMed] [Google Scholar]
  • 188.Versura P, Cellini M, Torreggiani A, Profazio V, Bernabini B, Caramazza R. Dryness symptoms, diagnostic protocol and therapeutic management: a report on 1,200 patients. Ophthalmic Res. 2001;33:221–7. doi: 10.1159/000055674. [DOI] [PubMed] [Google Scholar]
  • 189.Julio G, Lluch S, Pujol P, Merindano MD. Effects of tear hyperosmolarity on conjunctival cells in mild to moderate dry eye. Ophthalmic Physiol Opt. 2012;32:317–23. doi: 10.1111/j.1475-1313.2012.00915.x. [DOI] [PubMed] [Google Scholar]
  • 190.Horwath J, Ettinger K, Bachernegg M, Bodner E, Schmut O. Ocular Ferning test - effect of temperature and humidity on tear Ferning patterns. Ophthalmologica. 2001;215:102–7. doi: 10.1159/000050838. [DOI] [PubMed] [Google Scholar]
  • 191.Norn M. Quantitative tear ferning. Methodologic and experimental investigations. Acta Ophthalmol (Copenh) 1988;66:201–5. doi: 10.1111/j.1755-3768.1988.tb04012.x. [DOI] [PubMed] [Google Scholar]
  • 192.Jackson JA, Perrigin JA. Relationship of impression cytology and tear ferning to reports of dry eye. J Am Optom Assoc. 1999;70:187–92. [PubMed] [Google Scholar]
  • 193.Versura P, Profazio V, Cellini M, Torreggiani A, Caramazza R. Eye discomfort and air pollution. Ophthalmologica. 1999;213:103–9. doi: 10.1159/000027401. [DOI] [PubMed] [Google Scholar]
  • 194.Battaglia Parodi M, Giusto DD. Ocular fluid ferning test and fractals. Ophthalmic Res. 1993;25:307–13. doi: 10.1159/000267330. [DOI] [PubMed] [Google Scholar]
  • 195.Versura P, Frigato M, Cellini M, Mule R, Malavolta N, Campos EC. Diagnostic performance of tear function tests in Sjogren’s syndrome patients. Eye (Lond) 2007;21:229–37. doi: 10.1038/sj.eye.6702204. [DOI] [PubMed] [Google Scholar]
  • 196.Rolando M, Baldi F, Zingirian M. The effect of hyperosmolarity on tear mucus ferning. Fortschr Ophthalmol. 1986;83:644–6. [PubMed] [Google Scholar]
  • 197.Tatlipinar S, Gedik S, Irkec M, Orhan M, Erdener U. Ocular ferning during the menstrual cycle in healthy women. Eur J Ophthalmol. 2001;11:15–8. doi: 10.1177/112067210101100104. [DOI] [PubMed] [Google Scholar]
  • 198.Puderbach S, Stolze HH. Tear ferning and other lacrimal tests in normal persons of different ages. Int Ophthalmol. 1991;15:391–5. doi: 10.1007/BF00137950. [DOI] [PubMed] [Google Scholar]
  • 199.Felberg S, Cordeiro H, Sato EH, Martini Filho D, Nishiwaki-Dantas MC, Endo RM, et al. Reproducibility of the classification of ocular ferning patterns in Sjogren’s syndrome patients. Arq Bras Oftalmol. 2008;71:228–33. doi: 10.1590/s0004-27492008000200017. [DOI] [PubMed] [Google Scholar]
  • 200.Norn MS. Diagnosis of dry eye. In: Lemp MA, Marquardt R, editors. The Dry Eye: A Comprehensive Guide. New York: Springer Verlag; 1992. pp. 133–82. [Google Scholar]
  • 201.Wyon NM, Wyon DP. Measurement of acute response to draught in the eye. Acta Ophthalmol (Copenh) 1987;65:385–92. doi: 10.1111/j.1755-3768.1987.tb07011.x. [DOI] [PubMed] [Google Scholar]
  • 202.Kogbe O, Liotet S. An interesting use of the study of tear ferning patterns in contactology. Ophthalmologica. 1987;194:150–3. doi: 10.1159/000309753. [DOI] [PubMed] [Google Scholar]
  • 203.Altman PL, Dittmer DS. Man Biology Data Book. Bethesda, Maryland: Federation of Americal Societies of Experimental Biology; 1974. Physical properties and chemical composition of tears; pp. 2032–40. [Google Scholar]
  • 204.Carney LG, Hill RM. Human tear pH. Diurnal variations. Arch Ophthalmol. 1976;94:821–4. doi: 10.1001/archopht.1976.03910030405011. [DOI] [PubMed] [Google Scholar]
  • 205.Norn MS. Tear fluid pH in normals, contact lens wearers, and pathological cases. Acta Ophthalmol (Copenh) 1988;66:485–9. doi: 10.1111/j.1755-3768.1988.tb04368.x. [DOI] [PubMed] [Google Scholar]
  • 206.Carney LG, Hill RM. Human tear buffering capacity. Arch Ophthalmol. 1979;97:951–2. doi: 10.1001/archopht.1979.01020010509028. [DOI] [PubMed] [Google Scholar]
  • 207.Abelson MB, Udell IJ, Weston JH. Normal human tear pH by direct measurement. Arch Ophthalmol. 1981;99:301. doi: 10.1001/archopht.1981.03930010303017. [DOI] [PubMed] [Google Scholar]
  • 208.Fischer FH, Wiederholt M. Human precorneal tear film pH measured by microelectrodes. Graefes Arch Clin Exp Ophthalmol. 1982;218:168–70. doi: 10.1007/BF02215658. [DOI] [PubMed] [Google Scholar]
  • 209.McCarey BE, Wilson LA. pH, osmolarity and temperature effects on the water content of hydrogel contact lenses. Contact Intraocul Lens Med J. 1982;8:158–67. [PubMed] [Google Scholar]
  • 210.Coles WH, Jaros PA. Dynamics of ocular surface pH. Br J Ophthalmol. 1984;68:549–52. doi: 10.1136/bjo.68.8.549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Norn M. Tear pH after instillation of buffer in vivo. Acta Ophthalmol Suppl. 1985;173:32–4. doi: 10.1111/j.1755-3768.1985.tb06834.x. [DOI] [PubMed] [Google Scholar]
  • 212.Andres S, Garcia ML, Espina M, Valero J, Valls O. Tear pH, air pollution, and contact lenses. Am J Optom Physiol Opt. 1988;65:627–31. doi: 10.1097/00006324-198808000-00006. [DOI] [PubMed] [Google Scholar]
  • 213.Chen FS, Maurice DM. The pH in the precorneal tear film and under a contact lens measured with a fluorescent probe. Exp Eye Res. 1990;50:251–9. doi: 10.1016/0014-4835(90)90209-d. [DOI] [PubMed] [Google Scholar]
  • 214.Khurana AK, Chaudhary R, Ahluwalia BK, Gupta S. Tear film profile in dry eye. Acta Ophthalmol (Copenh) 1991;69:79–86. doi: 10.1111/j.1755-3768.1991.tb01997.x. [DOI] [PubMed] [Google Scholar]
  • 215.Yamada M, Mochizuki H, Kawai M, Yoshino M, Mashima Y. Fluorophotometric measurement of pH of human tears in vivo. Curr Eye Res. 1997;16:482–6. doi: 10.1076/ceyr.16.5.482.7050. [DOI] [PubMed] [Google Scholar]
  • 216.Janszky I, Vamosi P, Orszagh I, Berta A. Demonstration of increasing standard pH value of lacrimal fluid with increase of flow rate. Acta Ophthalmol Scand. 2001;79:180–3. doi: 10.1034/j.1600-0420.2001.079002180.x. [DOI] [PubMed] [Google Scholar]
  • 217.Mathers WD. Ocular evaporation in meibomian gland dysfunction and dry eye. Ophthalmology. 1993;100:347–51. doi: 10.1016/s0161-6420(93)31643-x. [DOI] [PubMed] [Google Scholar]
  • 218.Shimazaki J, Sakata M, Tsubota K. Ocular surface changes and discomfort in patients with meibomian gland dysfunction. Arch Ophthalmol. 1995;113:1266–70. doi: 10.1001/archopht.1995.01100100054027. [DOI] [PubMed] [Google Scholar]
  • 219.Mathers WD, Lane JA. Meibomian gland lipids, evaporation, and tear film stability. Adv Exp Med Biol. 1998;438:349–60. doi: 10.1007/978-1-4615-5359-5_50. [DOI] [PubMed] [Google Scholar]
  • 220.Khanal S, Tomlinson A, McFadyen A, Diaper C, Ramaesh K. Dry eye diagnosis. Invest Ophthalmol Vis Sci. 2008;49:1407–14. doi: 10.1167/iovs.07-0635. [DOI] [PubMed] [Google Scholar]
  • 221.Rolando M, Refojo MF, Kenyon KR. Increased tear evaporation in eyes with keratoconjunctivitis sicca. Arch Ophthalmol. 1983;101:557–8. doi: 10.1001/archopht.1983.01040010557003. [DOI] [PubMed] [Google Scholar]
  • 222.Mathers WD, Daley TE. Tear flow and evaporation in patients with and without dry eye. Ophthalmology. 1996;103:664–9. doi: 10.1016/s0161-6420(96)30637-4. [DOI] [PubMed] [Google Scholar]
  • 223.Hamano H, Hori M, Mitsunaga S. Application of an evaporimeter to the field of ophthalmology. J Jpn Contact Lens Soc. 1980;22:101–7. [Google Scholar]
  • 224.Mathers W. Evaporation from the ocular surface. Exp Eye Res. 2004;78:389–94. doi: 10.1016/s0014-4835(03)00199-4. [DOI] [PubMed] [Google Scholar]
  • 225.Herok GH, Mudgil P, Millar TJ. The effect of Meibomian lipids and tear proteins on evaporation rate under controlled in vitro conditions. Curr Eye Res. 2009;34:589–97. doi: 10.1080/02713680902972366. [DOI] [PubMed] [Google Scholar]
  • 226.Hamano H, Hori M, Mitsunaga S. Measurement of evaporation rate of water from the precorneal tear film and contact lenses. Contactology. 1981;25:8–14. [Google Scholar]
  • 227.Cedarstaff TH, Tomlinson A. A comparative study of tear evaporation rates and water content of soft contact lenses. Am J Optom Physiol Opt. 1983;60:167–74. doi: 10.1097/00006324-198303000-00004. [DOI] [PubMed] [Google Scholar]
  • 228.Rolando M, Refojo MF. Tear evaporimeter for measuring water evaporation rate from the tear film under controlled conditions in humans. Exp Eye Res. 1983;36:25–33. doi: 10.1016/0014-4835(83)90086-6. [DOI] [PubMed] [Google Scholar]
  • 229.Trees GR, Tomlinson A. Effect of artificial tear solutions and saline on tear film evaporation. Optom Vis Sci. 1990;67:886–90. doi: 10.1097/00006324-199012000-00002. [DOI] [PubMed] [Google Scholar]
  • 230.Tsubota K, Yamada M. Tear evaporation from the ocular surface. Invest Ophthalmol Vis Sci. 1992;33:2942–50. [PubMed] [Google Scholar]
  • 231.Mathers WD, Binarao G, Petroll M. Ocular water evaporation and the dry eye. A new measuring device. Cornea. 1993;12:335–40. doi: 10.1097/00003226-199307000-00010. [DOI] [PubMed] [Google Scholar]
  • 232.Goto E, Endo K, Suzuki A, Fujikura Y, Matsumoto Y, Tsubota K. Tear evaporation dynamics in normal subjects and subjects with obstructive meibomian gland dysfunction. Invest Ophthalmol Vis Sci. 2003;44:533–9. doi: 10.1167/iovs.02-0170. [DOI] [PubMed] [Google Scholar]
  • 233.Arciniega JC, Wojtowicz JC, Mohamed EM, McCulley JP. Changes in the evaporation rate of tear film after digital expression of meibomian glands in patients with and without dry eye. Cornea. 2011;30:843–7. doi: 10.1097/ICO.0b013e31820cd291. [DOI] [PubMed] [Google Scholar]
  • 234.Kojima T, Matsumoto Y, Ibrahim OM, Wakamatsu TH, Uchino M, Fukagawa K, et al. Effect of controlled adverse chamber environment exposure on tear functions in silicon hydrogel and hydrogel soft contact lens wearers. Invest Ophthalmol Vis Sci. 2011;52:8811–7. doi: 10.1167/iovs.10-6841. [DOI] [PubMed] [Google Scholar]
  • 235.Petznick A, Tan JH, Boo SK, Lee SY, Acharya UR, Tong L. Repeatability of a new method for measuring tear evaporation rates. Optom Vis Sci. 2013;90:366–71. doi: 10.1097/OPX.0b013e318288bdd1. [DOI] [PubMed] [Google Scholar]
  • 236.Thai LC, Tomlinson A, Doane MG. Effect of contact lens materials on tear physiology. Optom Vis Sci. 2004;81:194–204. doi: 10.1097/00006324-200403000-00012. [DOI] [PubMed] [Google Scholar]
  • 237.Guillon M, Maissa C. Contact lens wear affects tear film evaporation. Eye Contact Lens. 2008;34:326–30. doi: 10.1097/ICL.0b013e31818c5d00. [DOI] [PubMed] [Google Scholar]
  • 238.Hirayama M, Murat D, Liu Y, Kojima T, Kawakita T, Tsubota K. Efficacy of a novel moist cool air device in office workers with dry eye disease. Acta Ophthalmol. 2013;91:756–62. doi: 10.1111/j.1755-3768.2012.02485.x. [DOI] [PubMed] [Google Scholar]
  • 239.Rohit A, Ehrmann K, Naduvilath T, Willcox M, Stapleton F. Validating a new device for measuring tear evaporation rates. Ophthalmic Physiol Opt. 2014;34:53–62. doi: 10.1111/opo.12096. [DOI] [PubMed] [Google Scholar]
  • 240.Yeo S, Tan JH, Acharya UR, Sudarshan VK, Tong L. Longitudinal changes in tear evaporation rates after eyelid warming therapies in meibomian gland dysfunction. Invest Ophthalmol Vis Sci. 2016;57:1974–81. doi: 10.1167/iovs.16-19088. [DOI] [PubMed] [Google Scholar]
  • 241.Alghamdi WM, Markoulli M, Holden BA, Papas EB. Impact of duration of contact lens wear on the structure and function of the meibomian glands. Ophthalmic Physiol Opt. 2016;36:120–31. doi: 10.1111/opo.12278. [DOI] [PubMed] [Google Scholar]
  • 242.Peng CC, Cerretani C, Braun RJ, Radke CJ. Evaporation-driven instability of the precorneal tear film. Adv Colloid Interface Sci. 2014;206:250–64. doi: 10.1016/j.cis.2013.06.001. [DOI] [PubMed] [Google Scholar]
  • 243.Kimball SH, King-Smith PE, Nichols JJ. Evidence for the major contribution of evaporation to tear film thinning between blinks. Invest Ophthalmol Vis Sci. 2010;51:6294–7. doi: 10.1167/iovs.09-4772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Mathers WD, Lane JA, Zimmerman MB. Tear film changes associated with normal aging. Cornea. 1996;15:229–34. doi: 10.1097/00003226-199605000-00001. [DOI] [PubMed] [Google Scholar]
  • 245.Guillon M, Maissa C. Tear film evaporation–effect of age and gender. Cont Lens Anterior Eye. 2010;33:171–5. doi: 10.1016/j.clae.2010.03.002. [DOI] [PubMed] [Google Scholar]
  • 246.Tan JH, Ng EY, Acharya UR. Evaluation of tear evaporation from ocular surface by functional infrared thermography. Med Phys. 2010;37:6022–34. doi: 10.1118/1.3495540. [DOI] [PubMed] [Google Scholar]
  • 247.Craig JP, Tomlinson A. Age and gender effects on the normal tear film. Adv Exp Med Biol. 1998;438:411–5. doi: 10.1007/978-1-4615-5359-5_57. [DOI] [PubMed] [Google Scholar]
  • 248.Tsubota K, Nakamori K. Effects of ocular surface area and blink rate on tear dynamics. Arch Ophthalmol. 1995;113:155–8. doi: 10.1001/archopht.1995.01100020037025. [DOI] [PubMed] [Google Scholar]
  • 249.McCulley JP, Aronowicz JD, Uchiyama E, Shine WE, Butovich IA. Correlations in a change in aqueous tear evaporation with a change in relative humidity and the impact. Am J Ophthalmol. 2006;141:758–60. doi: 10.1016/j.ajo.2005.10.057. [DOI] [PubMed] [Google Scholar]
  • 250.Uchiyama E, Aronowicz JD, Butovich IA, McCulley JP. Increased evaporative rates in laboratory testing conditions simulating airplane cabin relative humidity: an important factor for dry eye syndrome. Eye Contact Lens. 2007;33:174–6. doi: 10.1097/01.icl.0000252881.04636.5e. [DOI] [PubMed] [Google Scholar]
  • 251.Abusharha AA, Pearce EI. The effect of low humidity on the human tear film. Cornea. 2013;32:429–34. doi: 10.1097/ICO.0b013e31826671ab. [DOI] [PubMed] [Google Scholar]
  • 252.Tomlinson A, Doane MG, McFadyen A. Inputs and outputs of the lacrimal system: review of production and evaporative loss. Ocul Surf. 2009;7:186–98. doi: 10.1016/s1542-0124(12)70186-6. [DOI] [PubMed] [Google Scholar]
  • 253.Rantamaki AH, Javanainen M, Vattulainen I, Holopainen JM. Do lipids retard the evaporation of the tear fluid? Invest Ophthalmol Vis Sci. 2012;53:6442–7. doi: 10.1167/iovs.12-10487. [DOI] [PubMed] [Google Scholar]
  • 254.Borchman D, Foulks GN, Yappert MC, Mathews J, Leake K, Bell J. Factors affecting evaporation rates of tear film components measured in vitro. Eye Contact Lens. 2009;35:32–7. doi: 10.1097/ICL.0b013e318193f4fc. [DOI] [PubMed] [Google Scholar]
  • 255.Millar TJ, Schuett BS. The real reason for having a meibomian lipid layer covering the outer surface of the tear film - a review. Exp Eye Res. 2015;137:125–38. doi: 10.1016/j.exer.2015.05.002. [DOI] [PubMed] [Google Scholar]
  • 256.Brown SI, Dervichian DG. The oils of the meibomian glands. Physical and surface characteristics. Arch Ophthalmol. 1969;82:537–40. doi: 10.1001/archopht.1969.00990020539019. [DOI] [PubMed] [Google Scholar]
  • 257.Cerretani CF, Ho NH, Radke CJ. Water-evaporation reduction by duplex films: application to the human tear film. Adv Colloid Interface Sci. 2013;197–198:33–57. doi: 10.1016/j.cis.2013.03.007. [DOI] [PubMed] [Google Scholar]
  • 258.Kulovesi P, Rantamaki AH, Holopainen JM. Surface properties of artificial tear film lipid layers: effects of wax esters. Invest Ophthalmol Vis Sci. 2014;55:4448–54. doi: 10.1167/iovs.14-14122. [DOI] [PubMed] [Google Scholar]
  • 259.La Mer VK, Healy TW. Evaporation of water: its retardation by monolayers: spreading a monomolecular film on the surface is a tested and economical means of reducing water loss. Science. 1965;148:36–42. doi: 10.1126/science.148.3666.36. [DOI] [PubMed] [Google Scholar]
  • 260.Paananen RO, Rantamaki AH, Holopainen JM. Antievaporative mechanism of wax esters: implications for the function of tear fluid. Langmuir. 2014;30:5897–902. doi: 10.1021/la501678t. [DOI] [PubMed] [Google Scholar]
  • 261.Rantamaki AH, Wiedmer SK, Holopainen JM. Melting points–the key to the anti-evaporative effect of the tear film wax esters. Invest Ophthalmol Vis Sci. 2013;54:5211–7. doi: 10.1167/iovs.13-12408. [DOI] [PubMed] [Google Scholar]
  • 262.Butovich IA, Lu H, McMahon A, Ketelson H, Senchyna M, Meadows D, et al. Biophysical and morphological evaluation of human normal and dry eye meibum using hot stage polarized light microscopy. Invest Ophthalmol Vis Sci. 2014;55:87–101. doi: 10.1167/iovs.13-13355. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Leiske D, Leiske C, Leiske D, Toney M, Senchyna M, Ketelson H, et al. Temperature-induced transitions in the structure and interfacial rheology of human meibum. Biophys J. 2012;102:369–76. doi: 10.1016/j.bpj.2011.12.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Lu H, Wojtowicz JC, Butovich IA. Differential scanning calorimetric evaluation of human meibomian gland secretions and model lipid mixtures: transition temperatures and cooperativity of melting. Chem Phys Lipids. 2013;170–171:55–64. doi: 10.1016/j.chemphyslip.2013.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Bhamla MS, Chai C, Rabiah NI, Frostad JM, Fuller GG. Instability and breakup of model tear films. Invest Ophthalmol Vis Sci. 2016;57:949–58. doi: 10.1167/iovs.15-18064. [DOI] [PubMed] [Google Scholar]
  • 266.McCulley JP, Shine W. A compositional based model for the tear film lipid layer. Trans Am Ophthalmol Soc. 1997;95:79–88. discussion -93. [PMC free article] [PubMed] [Google Scholar]
  • 267.Iwata S, Lemp MA, Holly FJ, Dohlman CH. Evaporation rate of water from the precorneal tear film and cornea in the rabbit. Invest Ophthalmol. 1969;8:613–9. [PubMed] [Google Scholar]
  • 268.de Souza GA, Godoy LM, Mann M. Identification of 491 proteins in the tear fluid proteome reveals a large number of proteases and protease inhibitors. Genome Biol. 2006;7:R72. doi: 10.1186/gb-2006-7-8-r72. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Green-Church KB, Nichols KK, Kleinholz NM, Zhang L, Nichols JJ. Investigation of the human tear film proteome using multiple proteomic approaches. Mol Vis. 2008;14:456–70. [PMC free article] [PubMed] [Google Scholar]
  • 270.Argueso P, Balaram M, Spurr-Michaud S, Keutmann HT, Dana MR, Gipson IK. Decreased levels of the goblet cell mucin MUC5AC in tears of patients with Sjogren syndrome. Invest Ophthalmol Vis Sci. 2002;43:1004–11. [PubMed] [Google Scholar]
  • 271.Spurr-Michaud S, Argueso P, Gipson I. Assay of mucins in human tear fluid. Exp Eye Res. 2007;84:939–50. doi: 10.1016/j.exer.2007.01.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 272.Butovich IA. On the lipid composition of human meibum and tears: comparative analysis of nonpolar lipids. Invest Ophthalmol Vis Sci. 2008;49:3779–89. doi: 10.1167/iovs.08-1889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Langmuir I, Schaefer VJ. Rates of evaporation of water through compressed monolayers on water. J Frankl Inst. 1943;235:119–62. [Google Scholar]
  • 274.Borchman D. Does the tear film lipid layer inhibit the rate of evaporation of tears? EC Ophthalmol. 2015;3:251–3. [Google Scholar]
  • 275.Coullerez G, Gorodyska G, Reimhult E, Textor M. Grandin HM self-assembled multifunctional polymers for biointerfaces. In: Knoll W, Advincula RC, editors. Functional Polymer Films. Vol. 2. Bognor Regis: John Wiley & Sons Ltd; 2012. pp. 855–905. [Google Scholar]
  • 276.Green-Church KB, Butovich I, Willcox M, Borchman D, Paulsen F, Barabino S, et al. The international workshop on meibomian gland dysfunction: report of the subcommittee on tear film lipids and lipid-protein interactions in health and disease. Invest Ophthalmol Vis Sci. 2011;52:1979–93. doi: 10.1167/iovs.10-6997d. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Tiffany JM. The lipid secretion of the meibomian glands. Adv Lipid Res. 1987;22:1–62. doi: 10.1016/b978-0-12-024922-0.50005-9. [DOI] [PubMed] [Google Scholar]
  • 278.Yokoi N, Yamada H, Mizukusa Y, Bron AJ, Tiffany JM, Kato T, et al. Rheology of tear film lipid layer spread in normal and aqueous tear-deficient dry eyes. Invest Ophthalmol Vis Sci. 2008;49:5319–24. doi: 10.1167/iovs.07-1407. [DOI] [PubMed] [Google Scholar]
  • 279.Rolando M, Valente C, Barabino S. New test to quantify lipid layer behavior in healthy subjects and patients with keratoconjunctivitis sicca. Cornea. 2008;27:866–70. doi: 10.1097/ICO.0b013e31816f630e. [DOI] [PubMed] [Google Scholar]
  • 280.Kulovesi P, Telenius J, Koivuniemi A, Brezesinski G, Rantamaki A, Viitala T, et al. Molecular organization of the tear fluid lipid layer. Biophys J. 2010;99:2559–67. doi: 10.1016/j.bpj.2010.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Kulovesi P, Telenius J, Koivuniemi A, Brezesinski G, Vattulainen I, Holopainen JM. The impact of lipid composition on the stability of the tear fluid lipid layer. Soft Matter. 2012;8:5826–34. [Google Scholar]
  • 282.Kunnen CM, Brown SH, Lazon de la Jara P, Holden BA, Blanksby SJ, Mitchell TW, et al. Influence of meibomian gland expression methods on human lipid analysis results. Ocul Surf. 2016;14:49–55. doi: 10.1016/j.jtos.2015.10.001. [DOI] [PubMed] [Google Scholar]
  • 283.Haworth KM, Nichols JJ, Thangavelu M, Sinnott LT, Nichols KK. Examination of human meibum collection and extraction techniques. Optom Vis Sci. 2011;88:525–33. doi: 10.1097/OPX.0b013e318214ac0f. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Pucker AD, Nichols JJ. Analysis of meibum and tear lipids. Ocul Surf. 2012;10:230–50. doi: 10.1016/j.jtos.2012.07.004. [DOI] [PubMed] [Google Scholar]
  • 285.Pucker AD, Haworth KM. The presence and significance of polar meibum and tear lipids. Ocul Surf. 2015;13:26–42. doi: 10.1016/j.jtos.2014.06.002. [DOI] [PubMed] [Google Scholar]
  • 286.Nagyova B, Tiffany JM. Components responsible for the surface tension of human tears. Curr Eye Res. 1999;19:4–11. doi: 10.1076/ceyr.19.1.4.5341. [DOI] [PubMed] [Google Scholar]
  • 287.Svitova TF, Lin MC. Dynamic interfacial properties of human tear-lipid films and their interactions with model-tear proteins in vitro. Adv Colloid Interface Sci. 2016;233:4–24. doi: 10.1016/j.cis.2015.12.009. [DOI] [PubMed] [Google Scholar]
  • 288.Borchman D, Foulks GN, Yappert MC, Tang D, Ho DV. Spectroscopic evaluation of human tear lipids. Chem Phys Lipids. 2007;147:87–102. doi: 10.1016/j.chemphyslip.2007.04.001. [DOI] [PubMed] [Google Scholar]
  • 289.Saville JT, Zhao Z, Willcox MD, Blanksby SJ, Mitchell TW. Detection and quantification of tear phospholipids and cholesterol in contact lens deposits: the effect of contact lens material and lens care solution. Invest Ophthalmol Vis Sci. 2010;51:2843–51. doi: 10.1167/iovs.09-4609. [DOI] [PubMed] [Google Scholar]
  • 290.Lam SM, Tong L, Duan X, Petznick A, Wenk MR, Shui G. Extensive characterization of human tear fluid collected using different techniques unravels the presence of novel lipid amphiphiles. J Lipid Res. 2014;55:289–98. doi: 10.1194/jlr.M044826. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Rohit A, Stapleton F, Brown SH, Mitchell TW, Willcox MD. Comparison of tear lipid profile among basal, reflex, and flush tear samples. Optom Vis Sci. 2014;91:1391–5. doi: 10.1097/OPX.0000000000000411. [DOI] [PubMed] [Google Scholar]
  • 292.Dean AW, Glasgow BJ. Mass spectrometric identification of phospholipids in human tears and tear lipocalin. Invest Ophthalmol Vis Sci. 2012;53:1773–82. doi: 10.1167/iovs.11-9419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Georgiev GA. Controversies regarding the role of polar lipids in human and animal tear film lipid layer. Ocul Surf. 2015;13:176–8. doi: 10.1016/j.jtos.2015.04.003. [DOI] [PubMed] [Google Scholar]
  • 294.Tiffany JM, Winter N, Bliss G. Tear film stability and tear surface tension. Curr Eye Res. 1989;8:507–15. doi: 10.3109/02713688909000031. [DOI] [PubMed] [Google Scholar]
  • 295.Tiffany JM. Tears in health and disease. Eye (Lond) 2003;17:923–6. doi: 10.1038/sj.eye.6700566. [DOI] [PubMed] [Google Scholar]
  • 296.Millar TJ, Tragoulias ST, Anderton PJ, Ball MS, Miano F, Dennis GR, et al. The surface activity of purified ocular mucin at the air-liquid interface and interactions with meibomian lipids. Cornea. 2006;25:91–100. doi: 10.1097/01.ico.0000164779.87795.3c. [DOI] [PubMed] [Google Scholar]
  • 297.Miano F, Calcara M, Millar TJ, Enea V. Insertion of tear proteins into a meibomian lipids film. Colloids Surf B Biointerfaces. 2005;44:49–55. doi: 10.1016/j.colsurfb.2005.05.011. [DOI] [PubMed] [Google Scholar]
  • 298.Millar TJ, Mudgil P, Butovich IA, Palaniappan CK. Adsorption of human tear lipocalin to human meibomian lipid films. Invest Ophthalmol Vis Sci. 2009;50:140–51. doi: 10.1167/iovs.08-2097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 299.Mudgil P, Torres M, Millar TJ. Adsorption of lysozyme to phospholipid and meibomian lipid monolayer films. Colloids Surf B Biointerfaces. 2006;48:128–37. doi: 10.1016/j.colsurfb.2006.01.017. [DOI] [PubMed] [Google Scholar]
  • 300.Palaniappan CK, Schutt BS, Brauer L, Schicht M, Millar TJ. Effects of keratin and lung surfactant proteins on the surface activity of meibomian lipids. Invest Ophthalmol Vis Sci. 2013;54:2571–81. doi: 10.1167/iovs.12-11084. [DOI] [PubMed] [Google Scholar]
  • 301.Leiske DL, Miller CE, Rosenfeld L, Cerretani C, Ayzner A, Lin B, et al. Molecular structure of interfacial human meibum films. Langmuir. 2012;28:11858–65. doi: 10.1021/la301321r. [DOI] [PubMed] [Google Scholar]
  • 302.Rosenfeld L, Cerretani C, Leiske DL, Toney MF, Radke CJ, Fuller GG. Structural and rheological properties of meibomian lipid. Invest Ophthalmol Vis Sci. 2013;54:2720–32. doi: 10.1167/iovs.12-10987. [DOI] [PubMed] [Google Scholar]
  • 303.Borchman D, Foulks GN, Yappert MC, Bell J, Wells E, Neravetla S, et al. Human meibum lipid conformation and thermodynamic changes with meibomian-gland dysfunction. Invest Ophthalmol Vis Sci. 2011;52:3805–17. doi: 10.1167/iovs.10-6514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.King-Smith PE, Bailey MD, Braun RJ. Four characteristics and a model of an effective tear film lipid layer (TFLL) Ocul Surf. 2013;11:236–45. doi: 10.1016/j.jtos.2013.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 305.Georgiev GA, Yokoi N, Ivanova S, Dimitov T, Andreev K, Krastev R, et al. Surface chemistry study of the interactions of hyaluronic acid and benzal-konium chloride with meibomian and corneal cell lipids. Soft Matter. 2013;9:10841–56. [Google Scholar]
  • 306.Leiske DL, Raju SR, Ketelson HA, Millar TJ, Fuller GG. The interfacial visco-elastic properties and structures of human and animal Meibomian lipids. Exp Eye Res. 2010;90:598–604. doi: 10.1016/j.exer.2010.02.004. [DOI] [PubMed] [Google Scholar]
  • 307.Tragoulias ST, Anderton PJ, Dennis GR, Miano F, Millar TJ. Surface pressure measurements of human tears and individual tear film components indicate that proteins are major contributors to the surface pressure. Cornea. 2005;24:189–200. doi: 10.1097/01.ico.0000138837.52694.37. [DOI] [PubMed] [Google Scholar]
  • 308.Millar TJ. A mechanism to explain the behaviour of spread films of meibomian lipids. Curr Eye Res. 2013;38:220–3. doi: 10.3109/02713683.2012.721045. [DOI] [PubMed] [Google Scholar]
  • 309.Butovich IA. Tear film lipids. Exp Eye Res. 2013;117:4–27. doi: 10.1016/j.exer.2013.05.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 310.Wizert A, Iskander DR, Cwiklik L. Organization of lipids in the tear film: a molecular-level view. PLoS One. 2014;9:e92461. doi: 10.1371/journal.pone.0092461. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 311.Cwiklik L. Tear film lipid layer: a molecular level view. Biochim Biophys Acta. 2016;1858:2421–30. doi: 10.1016/j.bbamem.2016.02.020. [DOI] [PubMed] [Google Scholar]
  • 312.Wei XE, Markoulli M, Millar TJ, Willcox MD, Zhao Z. Divalent cations in tears, and their influence on tear film stability in humans and rabbits. Invest Ophthalmol Vis Sci. 2012;53:3280–5. doi: 10.1167/iovs.12-9558. [DOI] [PubMed] [Google Scholar]
  • 313.Mudgil P, Millar TJ. Surfactant properties of human meibomian lipids. Invest Ophthalmol Vis Sci. 2011;52:1661–70. doi: 10.1167/iovs.10-5445. [DOI] [PubMed] [Google Scholar]
  • 314.Svitova TF, Lin MC. Tear lipids interfacial rheology: effect of lysozyme and lens care solutions. Optom Vis Sci. 2010;87:10–20. doi: 10.1097/OPX.0b013e3181c07908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Georgiev GA, Yokoi N, Koev K, Kutsarova E, Ivanova S, Kyumurkov A, et al. Surface chemistry study of the interactions of benzalkonium chloride with films of meibum, corneal cells lipids, and whole tears. Invest Ophthalmol Vis Sci. 2011;52:4645–54. doi: 10.1167/iovs.10-6271. [DOI] [PubMed] [Google Scholar]
  • 316.Georgiev GA, Yokoi N, Ivanova S, Krastev R, Lalchev Z. Surface chemistry study of the interactions of pharmaceutical ingredients with human meibum films. Invest Ophthalmol Vis Sci. 2012;53:4605–15. doi: 10.1167/iovs.12-9907. [DOI] [PubMed] [Google Scholar]
  • 317.Svitova TF, Lin MC. Lens-care-solution-induced alterations in dynamic interfacial properties of human tear-lipid films. Cont Lens Anterior Eye. 2014;37:368–76. doi: 10.1016/j.clae.2014.05.008. [DOI] [PubMed] [Google Scholar]
  • 318.Ivanova S, Tonchev V, Yokoi N, Yappert MC, Borchman D, Georgiev GA. Surface properties of squalene/meibum films and NMR confirmation of squalene in tears. Int J Mol Sci. 2015;16:21813–31. doi: 10.3390/ijms160921813. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Georgiev GA, Kutsarova E, Jordanova A, Krastev R, Lalchev Z. Interactions of Meibomian gland secretion with polar lipids in Langmuir monolayers. Colloids Surfaces B Biointerfaces. 2010;78:317–27. doi: 10.1016/j.colsurfb.2010.03.024. [DOI] [PubMed] [Google Scholar]
  • 320.Schuett BS, Millar TJ. An investigation of the likely role of (O-acyl) omega-hydroxy fatty acids in meibomian lipid films using (O-oleyl) omega-hydroxy palmitic acid as a model. Exp Eye Res. 2013;115:57–64. doi: 10.1016/j.exer.2013.06.016. [DOI] [PubMed] [Google Scholar]
  • 321.Millar TJ, King-Smith PE. Analysis of comparison of human meibomian lipid films and mixtures with cholesteryl esters in vitro films using high resolution color microscopy. Invest Ophthalmol Vis Sci. 2012;53:4710–9. doi: 10.1167/iovs.12-10022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Schuett BS, Millar TJ. Lipid component contributions to the surface activity of meibomian lipids. Invest Ophthalmol Vis Sci. 2012;53:7208–19. doi: 10.1167/iovs.12-10471. [DOI] [PubMed] [Google Scholar]
  • 323.Mochizuki H, Yamada M, Hatou S, Tsubota K. Turnover rate of tear-film lipid layer determined by fluorophotometry. Br J Ophthalmol. 2009;93:1535–8. doi: 10.1136/bjo.2008.156828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 324.Sharma A, Ruckenstein E. Mechanism of tear film rupture and its implications for contact lens tolerance. Am J Optom Physiol Opt. 1985;62:246–53. doi: 10.1097/00006324-198504000-00003. [DOI] [PubMed] [Google Scholar]
  • 325.Craig JP, Wang MT, Kim D, Lee JM. Exploring the Predisposition of the Asian Eye to Development of Dry Eye. Ocul Surf. 2016;14:385–92. doi: 10.1016/j.jtos.2016.03.002. [DOI] [PubMed] [Google Scholar]
  • 326.Gouveia SM, Tiffany JM. Human tear viscosity: an interactive role for proteins and lipids. Biochim Biophys Acta. 2005;1753:155–63. doi: 10.1016/j.bbapap.2005.08.023. [DOI] [PubMed] [Google Scholar]
  • 327.Pandit JC, Nagyova B, Bron AJ, Tiffany JM. Physical properties of stimulated and unstimulated tears. Exp Eye Res. 1999;68:247–53. doi: 10.1006/exer.1998.0600. [DOI] [PubMed] [Google Scholar]
  • 328.Tiffany JM. The viscosity of human tears. Int Ophthalmol. 1991;15:371–6. doi: 10.1007/BF00137947. [DOI] [PubMed] [Google Scholar]
  • 329.Knop N, Knop E. Meibomian glands. Part I: anatomy, embryology and histology of the Meibomian glands. Ophthalmologe. 2009;106:872–83. doi: 10.1007/s00347-009-2006-1. [DOI] [PubMed] [Google Scholar]
  • 330.Blackie CA, Korb DR. The diurnal secretory characteristics of individual meibomian glands. Cornea. 2010;29:34–8. doi: 10.1097/ICO.0b013e3181ac9fd0. [DOI] [PubMed] [Google Scholar]
  • 331.Chew CK, Jansweijer C, Tiffany JM, Dikstein S, Bron AJ. An instrument for quantifying meibomian lipid on the lid margin: the Meibometer. Curr Eye Res. 1993;12:247–54. doi: 10.3109/02713689308999470. [DOI] [PubMed] [Google Scholar]
  • 332.Chew CK, Hykin PG, Jansweijer C, Dikstein S, Tiffany JM, Bron AJ. The casual level of meibomian lipids in humans. Curr Eye Res. 1993;12:255–9. doi: 10.3109/02713689308999471. [DOI] [PubMed] [Google Scholar]
  • 333.Lam SM, Tong L, Reux B, Duan X, Petznick A, Yong SS, et al. Lipidomic analysis of human tear fluid reveals structure-specific lipid alterations in dry eye syndrome. J Lipid Res. 2014;55:299–306. doi: 10.1194/jlr.P041780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Butovich IA. The Meibomian puzzle: combining pieces together. Prog Retin Eye Res. 2009;28:483–98. doi: 10.1016/j.preteyeres.2009.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Butovich IA. Cholesteryl esters as a depot for very long chain fatty acids in human meibum. J Lipid Res. 2009;50:501–13. doi: 10.1194/jlr.M800426-JLR200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 336.Lam SM, Tong L, Yong SS, Li B, Chaurasia SS, Shui G, et al. Meibum lipid composition in Asians with dry eye disease. PLoS One. 2011;6:e24339. doi: 10.1371/journal.pone.0024339. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Butovich IA, Wojtowicz JC, Molai M. Human tear film and meibum. Very long chain wax esters and (O-acyl)-omega-hydroxy fatty acids of meibum. J Lipid Res. 2009;50:2471–85. doi: 10.1194/jlr.M900252-JLR200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 338.Shine WE, McCulley JP. The role of cholesterol in chronic blepharitis. Invest Ophthalmol Vis Sci. 1991;32:2272–80. [PubMed] [Google Scholar]
  • 339.Borchman D, Yappert MC, Milliner SE, Duran D, Cox GW, Smith RJ, et al. 13C and 1H NMR ester region resonance assignments and the composition of human infant and child meibum. Exp Eye Res. 2013;112:151–9. doi: 10.1016/j.exer.2013.04.017. [DOI] [PubMed] [Google Scholar]
  • 340.Sullivan BD, Evans JE, Dana MR, Sullivan DA. Influence of aging on the polar and neutral lipid profiles in human meibomian gland secretions. Arch Ophthalmol. 2006;124:1286–92. doi: 10.1001/archopht.124.9.1286. [DOI] [PubMed] [Google Scholar]
  • 341.Brown SH, Kunnen CM, Papas EB, Lazon de la Jara P, Willcox MD, Blanksby SJ, et al. Intersubject and Interday Variability in Human Tear and Meibum Lipidomes: A Pilot Study. Ocul Surf. 2016;14:43–8. doi: 10.1016/j.jtos.2015.08.005. [DOI] [PubMed] [Google Scholar]
  • 342.Borchman D, Yappert MC, Milliner SE, Smith RJ, Bhola R. Confirmation of the presence of squalene in human eyelid lipid by heteronuclear single quantum correlation spectroscopy. Lipids. 2013;48:1269–77. doi: 10.1007/s11745-013-3844-9. [DOI] [PubMed] [Google Scholar]
  • 343.Kelly GS. Squalene and its potential clinical uses. Altern Med Rev. 1999;4:29–36. [PubMed] [Google Scholar]
  • 344.Lam SM, Tong L, Duan X, Acharya UR, Tan JH, Petznick A, et al. Longitudinal changes in tear fluid lipidome brought about by eyelid-warming treatment in a cohort of meibomian gland dysfunction. J Lipid Res. 2014;55:1959–69. doi: 10.1194/jlr.P051185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 345.Lewis BA, Engelman DM. Lipid bilayer thickness varies linearly with acyl chain length in fluid phosphatidylcholine vesicles. J Mol Biol. 1983;166:211–7. doi: 10.1016/s0022-2836(83)80007-2. [DOI] [PubMed] [Google Scholar]
  • 346.Borchman D, Foulks GN, Yappert MC, Milliner SE. Differences in human meibum lipid composition with meibomian gland dysfunction using NMR and principal component analysis. Invest Ophthalmol Vis Sci. 2012;53:337–47. doi: 10.1167/iovs.11-8551. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 347.Shrestha RK, Borchman D, Foulks GN, Yappert MC, Milliner SE. Analysis of the composition of lipid in human meibum from normal infants, children, adolescents, adults, and adults with meibomian gland dysfunction using (1) H-NMR spectroscopy. Invest Ophthalmol Vis Sci. 2011;52:7350–8. doi: 10.1167/iovs.11-7391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 348.Borchman D, Foulks GN, Yappert MC, Milliner SE. Changes in human meibum lipid composition with age using nuclear magnetic resonance spectroscopy. Invest Ophthalmol Vis Sci. 2012;53:475–82. doi: 10.1167/iovs.11-8341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 349.Borchman D, Foulks GN, Yappert MC, Kakar S, Podoll N, Rychwalski P, et al. Physical changes in human meibum with age as measured by infrared spectroscopy. Ophthalmic Res. 2010;44:34–42. doi: 10.1159/000283606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 350.Borchman D, Yappert MC, Foulks GN. Changes in human meibum lipid with meibomian gland dysfunction using principal component analysis. Exp Eye Res. 2010;91:246–56. doi: 10.1016/j.exer.2010.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 351.Saville JT, Zhao Z, Willcox MD, Ariyavidana MA, Blanksby SJ, Mitchell TW. Identification of phospholipids in human meibum by nano-electrospray ionisation tandem mass spectrometry. Exp Eye Res. 2011;92:238–40. doi: 10.1016/j.exer.2010.12.012. [DOI] [PubMed] [Google Scholar]
  • 352.Rantamaki AH, Seppanen-Laakso T, Oresic M, Jauhiainen M, Holopainen JM. Human tear fluid lipidome: from composition to function. PLoS One. 2011;6:e19553. doi: 10.1371/journal.pone.0019553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 353.Strott CA, Higashi Y. Cholesterol sulfate in human physiology: what’s it all about? J Lipid Res. 2003;44:1268–78. doi: 10.1194/jlr.R300005-JLR200. [DOI] [PubMed] [Google Scholar]
  • 354.Hattrup CL, Gendler SJ. Structure and function of the cell surface (tethered) mucins. Annu Rev Physiol. 2008;70:431–57. doi: 10.1146/annurev.physiol.70.113006.100659. [DOI] [PubMed] [Google Scholar]
  • 355.Thornton DJ, Rousseau K, McGuckin MA. Structure and function of the polymeric mucins in airways mucus. Annu Rev Physiol. 2008;70:459–86. doi: 10.1146/annurev.physiol.70.113006.100702. [DOI] [PubMed] [Google Scholar]
  • 356.Corfield AP. Mucins: a biologically relevant glycan barrier in mucosal protection. Biochim Biophys Acta. 2015;1850:236–52. doi: 10.1016/j.bbagen.2014.05.003. [DOI] [PubMed] [Google Scholar]
  • 357.Carraway KL, Perez A, Idris N, Jepson S, Arango M, Komatsu M, et al. Muc4/sialomucin complex, the intramembrane ErbB2 ligand, in cancer and epithelia: to protect and to survive. Prog Nucleic Acid Res Mol Biol. 2002;71:149–85. doi: 10.1016/s0079-6603(02)71043-x. [DOI] [PubMed] [Google Scholar]
  • 358.Dartt DA. Control of mucin production by ocular surface epithelial cells. Exp Eye Res. 2004;78:173–85. doi: 10.1016/j.exer.2003.10.005. [DOI] [PubMed] [Google Scholar]
  • 359.Hodges RR, Dartt DA. Tear film mucins: front line defenders of the ocular surface; comparison with airway and gastrointestinal tract mucins. Exp Eye Res. 2013;117:62–78. doi: 10.1016/j.exer.2013.07.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 360.Stephens DN, McNamara NA. Altered mucin and glycoprotein expression in dry eye disease. Optom Vis Sci. 2015;92:931–8. doi: 10.1097/OPX.0000000000000664. [DOI] [PubMed] [Google Scholar]
  • 361.Argueso P. Glycobiology of the ocular surface: mucins and lectins. Jpn J Ophthalmol. 2013;57:150–5. doi: 10.1007/s10384-012-0228-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 362.Mantelli F, Argueso P. Functions of ocular surface mucins in health and disease. Curr Opin Allergy Clin Immunol. 2008;8:477–83. doi: 10.1097/ACI.0b013e32830e6b04. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 363.Paulsen F, Langer G, Hoffmann W, Berry M. Human lacrimal gland mucins. Cell Tissue Res. 2004;316:167–77. doi: 10.1007/s00441-004-0877-7. [DOI] [PubMed] [Google Scholar]
  • 364.Yu DF, Chen Y, Han JM, Zhang H, Chen XP, Zou WJ, et al. MUC19 expression in human ocular surface and lacrimal gland and its alteration in Sjogren syndrome patients. Exp Eye Res. 2008;86:403–11. doi: 10.1016/j.exer.2007.11.013. [DOI] [PubMed] [Google Scholar]
  • 365.Moore JE, Vasey GT, Dartt DA, McGilligan VE, Atkinson SD, Grills C, et al. Effect of tear hyperosmolarity and signs of clinical ocular surface pathology upon conjunctival goblet cell function in the human ocular surface. Invest Ophthalmol Vis Sci. 2011;52:6174–80. doi: 10.1167/iovs.10-7022. [DOI] [PubMed] [Google Scholar]
  • 366.Zhang J, Yan X, Li H. Analysis of the correlations of mucins, inflammatory markers, and clinical tests in dry eye. Cornea. 2013;32:928–32. doi: 10.1097/ICO.0b013e3182801622. [DOI] [PubMed] [Google Scholar]
  • 367.Shimazaki-Den S, Dogru M, Higa K, Shimazaki J. Symptoms, visual function, and mucin expression of eyes with tear film instability. Cornea. 2013;32:1211–8. doi: 10.1097/ICO.0b013e318295a2a5. [DOI] [PubMed] [Google Scholar]
  • 368.Uchino Y, Uchino M, Yokoi N, Dogru M, Kawashima M, Okada N, et al. Alteration of tear mucin 5AC in office workers using visual display terminals: The Osaka Study. JAMA Ophthalmol. 2014;132:985–92. doi: 10.1001/jamaophthalmol.2014.1008. [DOI] [PubMed] [Google Scholar]
  • 369.Dogru M, Matsumoto Y, Okada N, Igarashi A, Fukagawa K, Shimazaki J, et al. Alterations of the ocular surface epithelial MUC16 and goblet cell MUC5AC in patients with atopic keratoconjunctivitis. Allergy. 2008;63:1324–34. doi: 10.1111/j.1398-9995.2008.01781.x. [DOI] [PubMed] [Google Scholar]
  • 370.Corrales RM, Calonge M, Herreras JM, Saez V, Chaves FJ. Human epithelium from conjunctival impression cytology expresses MUC7 mucin gene. Cornea. 2003;22:665–71. doi: 10.1097/00003226-200310000-00010. [DOI] [PubMed] [Google Scholar]
  • 371.Jumblatt MM, McKenzie RW, Steele PS, Emberts CG, Jumblatt JE. MUC7 expression in the human lacrimal gland and conjunctiva. Cornea. 2003;22:41–5. doi: 10.1097/00003226-200301000-00010. [DOI] [PubMed] [Google Scholar]
  • 372.Gomes GP, Assis MA, Fonseca JS, de Souza PE, Zenobio EG, Oliveira DD, et al. Genetic polymorphism of MUC7 in individuals with aggressive or chronic periodontitis. J Oral Sci. 2011;53:445–9. doi: 10.2334/josnusd.53.445. [DOI] [PubMed] [Google Scholar]
  • 373.Gururaja TL, Ramasubbu N, Venugopalan P, Reddy MS, Ramalingam K, Levine MJ. Structural features of the human salivary mucin, MUC7. Glycoconj J. 1998;15:457–67. doi: 10.1023/a:1006978818555. [DOI] [PubMed] [Google Scholar]
  • 374.Corrales RM, Narayanan S, Fernandez I, Mayo A, Galarreta DJ, Fuentes-Paez G, et al. Ocular mucin gene expression levels as biomarkers for the diagnosis of dry eye syndrome. Invest Ophthalmol Vis Sci. 2011;52:8363–9. doi: 10.1167/iovs.11-7655. [DOI] [PubMed] [Google Scholar]
  • 375.Pflugfelder SC, Liu Z, Monroy D, Li DQ, Carvajal ME, Price-Schiavi SA, et al. Detection of sialomucin complex (MUC4) in human ocular surface epithelium and tear fluid. Invest Ophthalmol Vis Sci. 2000;41:1316–26. [PubMed] [Google Scholar]
  • 376.Mantelli F, Schaffer L, Dana R, Head SR, Argueso P. Glycogene expression in conjunctiva of patients with dry eye: downregulation of Notch signaling. Invest Ophthalmol Vis Sci. 2009;50:2666–72. doi: 10.1167/iovs.08-2734. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 377.Woodward AM, Argueso P. Expression analysis of the transmembrane mucin MUC20 in human corneal and conjunctival epithelia. Invest Ophthalmol Vis Sci. 2014;55:6132–8. doi: 10.1167/iovs.14-15269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 378.Singh PK, Hollingsworth MA. Cell surface-associated mucins in signal transduction. Trends Cell Biol. 2006;16:467–76. doi: 10.1016/j.tcb.2006.07.006. [DOI] [PubMed] [Google Scholar]
  • 379.Shen Q, Rahn JJ, Zhang J, Gunasekera N, Sun X, Shaw AR, et al. MUC1 initiates Src-CrkL-Rac1/Cdc42-mediated actin cytoskeletal protrusive motility after ligating intercellular adhesion molecule-1. Mol Cancer Res. 2008;6:555–67. doi: 10.1158/1541-7786.MCR-07-2033. [DOI] [PubMed] [Google Scholar]
  • 380.Gipson IK, Spurr-Michaud S, Tisdale A, Menon BB. Comparison of the transmembrane mucins MUC1 and MUC16 in epithelial barrier function. PLoS One. 2014;9:e100393. doi: 10.1371/journal.pone.0100393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.Corrales RM, Galarreta D, Herreras JM, Saez V, Arranz I, Gonzalez MJ, et al. Conjunctival mucin mRNA expression in contact lens wear. Optom Vis Sci. 2009;86:1051–8. doi: 10.1097/OPX.0b013e3181b4f02e. [DOI] [PubMed] [Google Scholar]
  • 382.Argueso P, Guzman-Aranguez A, Mantelli F, Cao Z, Ricciuto J, Panjwani N. Association of cell surface mucins with galectin-3 contributes to the ocular surface epithelial barrier. J Biol Chem. 2009;284:23037–45. doi: 10.1074/jbc.M109.033332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Uchino Y, Mauris J, Woodward AM, Dieckow J, Amparo F, Dana R, et al. Alteration of galectin-3 in tears of patients with dry eye disease. Am J Ophthalmol. 2015;159:1027–35. doi: 10.1016/j.ajo.2015.02.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 384.Gipson IK, Spurr-Michaud SJ, Senchyna M, Ritter R, 3rd, Schaumberg D. Comparison of mucin levels at the ocular surface of postmenopausal women with and without a history of dry eye. Cornea. 2011;30:1346–52. doi: 10.1097/ICO.0b013e31820d852a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 385.Caffery B, Heynen ML, Joyce E, Jones L, Ritter R, 3rd, Senchyna M. MUC1 expression in Sjogren’s syndrome, KCS, and control subjects. Mol Vis. 2010;16:1720–7. [PMC free article] [PubMed] [Google Scholar]
  • 386.Caffery B, Joyce E, Heynen ML, Jones L, Ritter R, 3rd, Gamache DA, et al. MUC16 expression in Sjogren’s syndrome, KCS, and control subjects. Mol Vis. 2008;14:2547–55. [PMC free article] [PubMed] [Google Scholar]
  • 387.Albertsmeyer AC, Kakkassery V, Spurr-Michaud S, Beeks O, Gipson IK. Effect of pro-inflammatory mediators on membrane-associated mucins expressed by human ocular surface epithelial cells. Exp Eye Res. 2010;90:444–51. doi: 10.1016/j.exer.2009.12.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 388.Li YY, Chang JW, Hsieh LL, Yeh KY. Neutralization of interleukin (IL)-10 released by monocytes/macrophages enhances the up-regulatory effect of monocyte/macrophage-derived IL-6 on expressions of IL-6 and MUC1, and migration in HT-29 colon cancer cells. Cell Immunol. 2010;265:164–71. doi: 10.1016/j.cellimm.2010.07.014. [DOI] [PubMed] [Google Scholar]
  • 389.Paulsen F, Jager K, Worlitzsch D, Brauer L, Schulze U, Schafer G, et al. Regulation of MUC16 by inflammatory mediators in ocular surface epithelial cell lines. Ann Anat. 2008;190:59–70. doi: 10.1016/j.aanat.2007.05.001. [DOI] [PubMed] [Google Scholar]
  • 390.Cermak JM, Krenzer KL, Sullivan RM, Dana MR, Sullivan DA. Is complete androgen insensitivity syndrome associated with alterations in the meibomian gland and ocular surface? Cornea. 2003;22:516–21. doi: 10.1097/00003226-200308000-00006. [DOI] [PubMed] [Google Scholar]
  • 391.Argueso P, Tisdale A, Mandel U, Letko E, Foster CS, Gipson IK. The cell-layer-and cell-type-specific distribution of GalNAc-transferases in the ocular surface epithelia is altered during keratinization. Invest Ophthalmol Vis Sci. 2003;44:86–92. doi: 10.1167/iovs.02-0181. [DOI] [PubMed] [Google Scholar]
  • 392.Argueso P, Sumiyoshi M. Characterization of a carbohydrate epitope defined by the monoclonal antibody H185: sialic acid O-acetylation on epithelial cell-surface mucins. Glycobiology. 2006;16:1219–28. doi: 10.1093/glycob/cwl041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 393.Danjo Y, Watanabe H, Tisdale AS, George M, Tsumura T, Abelson MB, et al. Alteration of mucin in human conjunctival epithelia in dry eye. Invest Ophthalmol Vis Sci. 1998;39:2602–9. [PubMed] [Google Scholar]
  • 394.Hayashi Y, Kao WW, Kohno N, Nishihara-Hayashi M, Shiraishi A, Uno T, et al. Expression patterns of sialylated epitope recognized by KL-6 monoclonal antibody in ocular surface epithelium of normals and dry eye patients. Invest Ophthalmol Vis Sci. 2004;45:2212–7. doi: 10.1167/iovs.03-0988. [DOI] [PubMed] [Google Scholar]
  • 395.Nakamura Y, Yokoi N, Tokushige H, Kinoshita S. Sialic Acid in human tear fluid decreases in dry eye. Jpn J Ophthalmol. 2004;48:519–23. doi: 10.1007/s10384-004-0111-x. [DOI] [PubMed] [Google Scholar]
  • 396.Vijmasi T, Chen FY, Chen YT, Gallup M, McNamara N. Topical administration of interleukin-1 receptor antagonist as a therapy for aqueous-deficient dry eye in autoimmune disease. Mol Vis. 2013;19:1957–65. [PMC free article] [PubMed] [Google Scholar]
  • 397.Schmidt TA, Sullivan DA, Knop E, Richards SM, Knop N, Liu S, et al. Transcription, translation, and function of lubricin, a boundary lubricant, at the ocular surface. JAMA Ophthalmol. 2013;131:766–76. doi: 10.1001/jamaophthalmol.2013.2385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 398.Morrison S, Sullivan DA, Sullivan BD, Sheardown H, Schmidt TA. Dose-dependent and synergistic effects of proteoglycan 4 on boundary lubrication at a human cornea-polydimethylsiloxane biointerface. Eye Contact Lens. 2012;38:27–35. doi: 10.1097/ICL.0b013e31823f7041. [DOI] [PubMed] [Google Scholar]
  • 399.Samsom M, Chan A, Iwabuchi Y, Subbaraman L, Jones L, Schmidt TA. In vitro friction testing of contact lenses and human ocular tissues: Effect of proteoglycan 4 (PRG4) Tribol Int. 2015;89:27–33. [Google Scholar]
  • 400.Samsom ML, Morrison S, Masala N, Sullivan BD, Sullivan DA, Sheardown H, et al. Characterization of full-length recombinant human Proteoglycan 4 as an ocular surface boundary lubricant. Exp Eye Res. 2014;127:14–9. doi: 10.1016/j.exer.2014.06.015. [DOI] [PubMed] [Google Scholar]
  • 401.Lambiase A, Sullivan BD, Schmidt TA, Sullivan DA, Jay GD, Truitt ER, 3rd, et al. A Two-Week, Randomized, Double-masked Study to Evaluate Safety and Efficacy of Lubricin (150 mug/mL) Eye Drops Versus Sodium Hyaluronate (HA) 0. 18% Eye Drops (Vismed(R)) in Patients with Moderate Dry Eye Disease. Ocul Surf. 2017;15:77–87. doi: 10.1016/j.jtos.2016.08.004. [DOI] [PubMed] [Google Scholar]
  • 402.Schulz BL, Oxley D, Packer NH, Karlsson NG. Identification of two highly sialylated human tear-fluid DMBT1 isoforms: the major high-molecular-mass glycoproteins in human tears. Biochem J. 2002;366:511–20. doi: 10.1042/BJ20011876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 403.Prakobphol A, Xu F, Hoang VM, Larsson T, Bergstrom J, Johansson I, et al. Salivary agglutinin, which binds Streptococcus mutans and Helicobacter pylori, is the lung scavenger receptor cysteine-rich protein gp-340. J Biol Chem. 2000;275:39860–6. doi: 10.1074/jbc.M006928200. [DOI] [PubMed] [Google Scholar]
  • 404.Leito JT, Ligtenberg AJ, van Houdt M, van den Berg TK, Wouters D. The bacteria binding glycoprotein salivary agglutinin (SAG/gp340) activates complement via the lectin pathway. Mol Immunol. 2011;49:185–90. doi: 10.1016/j.molimm.2011.08.010. [DOI] [PubMed] [Google Scholar]
  • 405.Karnati R, Laurie DE, Laurie GW. Lacritin and the tear proteome as natural replacement therapy for dry eye. Exp Eye Res. 2013;117:39–52. doi: 10.1016/j.exer.2013.05.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 406.Semba RD, Enghild JJ, Venkatraman V, Dyrlund TF, Van Eyk JE. The Human Eye Proteome Project: perspectives on an emerging proteome. Proteomics. 2013;13:2500–11. doi: 10.1002/pmic.201300075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 407.Zhou L, Beuerman RW, Chan CM, Zhao SZ, Li XR, Yang H, et al. Identification of tear fluid biomarkers in dry eye syndrome using iTRAQ quantitative proteomics. J Proteom Res. 2009;8:4889–905. doi: 10.1021/pr900686s. [DOI] [PubMed] [Google Scholar]
  • 408.Nichols JJ, Green-Church KB. Mass spectrometry-based proteomic analyses in contact lens-related dry eye. Cornea. 2009;28:1109–17. doi: 10.1097/ICO.0b013e3181a2ad81. [DOI] [PubMed] [Google Scholar]
  • 409.Soria J, Duran JA, Etxebarria J, Merayo J, Gonzalez N, Reigada R, et al. Tear proteome and protein network analyses reveal a novel pentamarker panel for tear film characterization in dry eye and meibomian gland dysfunction. J Proteom. 2013;78:94–112. doi: 10.1016/j.jprot.2012.11.017. [DOI] [PubMed] [Google Scholar]
  • 410.Boehm N, Funke S, Wiegand M, Wehrwein N, Pfeiffer N, Grus FH. Alterations in the tear proteome of dry eye patients–a matter of the clinical phenotype. Invest Ophthalmol Vis Sci. 2013;54:2385–92. doi: 10.1167/iovs.11-8751. [DOI] [PubMed] [Google Scholar]
  • 411.Li B, Sheng M, Li J, Yan G, Lin A, Li M, et al. Tear proteomic analysis of Sjogren syndrome patients with dry eye syndrome by two-dimensional-nano-liquid chromatography coupled with tandem mass spectrometry. Sci Rep. 2014;4:5772. doi: 10.1038/srep05772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 412.Tong L, Zhou XY, Jylha A, Aapola U, Liu DN, Koh SK, et al. Quantitation of 47 human tear proteins using high resolution multiple reaction monitoring (HR-MRM) based-mass spectrometry. J Proteom. 2015;115:36–48. doi: 10.1016/j.jprot.2014.12.002. [DOI] [PubMed] [Google Scholar]
  • 413.Tsai PS, Evans JE, Green KM, Sullivan RM, Schaumberg DA, Richards SM, et al. Proteomic analysis of human meibomian gland secretions. Br J Ophthalmol. 2006;90:372–7. doi: 10.1136/bjo.2005.080846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 414.Perumal N, Funke S, Pfeiffer N, Grus FH. Proteomics analysis of human tears from aqueous-deficient and evaporative dry eye patients. Sci Rep. 2016;6:29629. doi: 10.1038/srep29629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 415.Acera A, Rocha G, Vecino E, Lema I, Duran JA. Inflammatory markers in the tears of patients with ocular surface disease. Ophthalmic Res. 2008;40:315–21. doi: 10.1159/000150445. [DOI] [PubMed] [Google Scholar]
  • 416.Seifert K, Gandia NC, Wilburn JK, Bower KS, Sia RK, Ryan DS, et al. Tear lacritin levels by age, sex, and time of day in healthy adults. Invest Ophthalmol Vis Sci. 2012;53:6610–6. doi: 10.1167/iovs.11-8729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 417.Enriquez-de-Salamanca A, Castellanos E, Stern ME, Fernandez I, Carreno E, Garcia-Vazquez C, et al. Tear cytokine and chemokine analysis and clinical correlations in evaporative-type dry eye disease. Mol Vis. 2010;16:862–73. [PMC free article] [PubMed] [Google Scholar]
  • 418.Acera A, Vecino E, Duran JA. Tear MMP-9 levels as a marker of ocular surface inflammation in conjunctivochalasis. Invest Ophthalmol Vis Sci. 2013;54:8285–91. doi: 10.1167/iovs.13-12235. [DOI] [PubMed] [Google Scholar]
  • 419.Lopez-Miguel A, Teson M, Martin-Montanez V, Enriquez-de-Salamanca A, Stern ME, Calonge M, et al. Dry eye exacerbation in patients exposed to desiccating stress under controlled environmental conditions. Am J Ophthalmol. 2014;157:788–98. doi: 10.1016/j.ajo.2014.01.001. [DOI] [PubMed] [Google Scholar]
  • 420.Zheng X, De Paiva CS, Rao K, Li DQ, Farley WJ, Stern M, et al. Evaluation of the transforming growth factor-beta activity in normal and dry eye human tears by CCL-185 cell bioassay. Cornea. 2010;29:1048–54. doi: 10.1097/ICO.0b013e3181cf98ff. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 421.Chan TC, Ye C, Chan KP, Chu KO, Jhanji V. Evaluation of point-of-care test for elevated tear matrix metalloproteinase 9 in post-LASIK dry eyes. Br J Ophthalmol. 2016;100:1188–91. doi: 10.1136/bjophthalmol-2015-307607. [DOI] [PubMed] [Google Scholar]
  • 422.Gonzalez N, Iloro I, Duran JA, Elortza F, Suarez T. Evaluation of inter-day and inter-individual variability of tear peptide/protein profiles by MALDI-TOF MS analyses. Mol Vis. 2012;18:1572–82. [PMC free article] [PubMed] [Google Scholar]
  • 423.Boehm N, Riechardt AI, Wiegand M, Pfeiffer N, Grus FH. Proinflammatory cytokine profiling of tears from dry eye patients by means of antibody microarrays. Invest Ophthalmol Vis Sci. 2011;52:7725–30. doi: 10.1167/iovs.11-7266. [DOI] [PubMed] [Google Scholar]
  • 424.Choi W, Li Z, Oh HJ, Im SK, Lee SH, Park SH, et al. Expression of CCR5 and its ligands CCL3, -4, and -5 in the tear film and ocular surface of patients with dry eye disease. Curr Eye Res. 2012;37:12–7. doi: 10.3109/02713683.2011.622852. [DOI] [PubMed] [Google Scholar]
  • 425.Lam H, Bleiden L, de Paiva CS, Farley W, Stern ME, Pflugfelder SC. Tear cytokine profiles in dysfunctional tear syndrome. Am J Ophthalmol. 2009;147:198–205. doi: 10.1016/j.ajo.2008.08.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 426.Lambiase A, Micera A, Sacchetti M, Cortes M, Mantelli F, Bonini S. Alterations of tear neuromediators in dry eye disease. Arch Ophthalmol. 2011;129:981–6. doi: 10.1001/archophthalmol.2011.200. [DOI] [PubMed] [Google Scholar]
  • 427.Massingale ML, Li X, Vallabhajosyula M, Chen D, Wei Y, Asbell PA. Analysis of inflammatory cytokines in the tears of dry eye patients. Cornea. 2009;28:1023–7. doi: 10.1097/ICO.0b013e3181a16578. [DOI] [PubMed] [Google Scholar]
  • 428.McKown RL, Wang N, Raab RW, Karnati R, Zhang Y, Williams PB, et al. Lacritin and other new proteins of the lacrimal functional unit. Exp Eye Res. 2009;88:848–58. doi: 10.1016/j.exer.2008.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 429.Na KS, Mok JW, Kim JY, Rho CR, Joo CK. Correlations between tear cytokines, chemokines, and soluble receptors and clinical severity of dry eye disease. Invest Ophthalmol Vis Sci. 2012;53:5443–50. doi: 10.1167/iovs.11-9417. [DOI] [PubMed] [Google Scholar]
  • 430.Ohashi Y, Ishida R, Kojima T, Goto E, Matsumoto Y, Watanabe K, et al. Abnormal protein profiles in tears with dry eye syndrome. Am J Ophthalmol. 2003;136:291–9. doi: 10.1016/s0002-9394(03)00203-4. [DOI] [PubMed] [Google Scholar]
  • 431.Rao K, Farley WJ, Pflugfelder SC. Association between high tear epidermal growth factor levels and corneal subepithelial fibrosis in dry eye conditions. Invest Ophthalmol Vis Sci. 2010;51:844–9. doi: 10.1167/iovs.09-3875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 432.Yoon KC, Jeong IY, Park YG, Yang SY. Interleukin-6 and tumor necrosis factor-alpha levels in tears of patients with dry eye syndrome. Cornea. 2007;26:431–7. doi: 10.1097/ICO.0b013e31803dcda2. [DOI] [PubMed] [Google Scholar]
  • 433.Solomon A, Dursun D, Liu Z, Xie Y, Macri A, Pflugfelder SC. Pro- and anti-inflammatory forms of interleukin-1 in the tear fluid and conjunctiva of patients with dry-eye disease. Invest Ophthalmol Vis Sci. 2001;42:2283–92. [PubMed] [Google Scholar]
  • 434.Virtanen T, Konttinen YT, Honkanen N, Harkonen M, Tervo T. Tear fluid plasmin activity of dry eye patients with Sjogren’s syndrome. Acta Ophthalmol Scand. 1997;75:137–41. doi: 10.1111/j.1600-0420.1997.tb00109.x. [DOI] [PubMed] [Google Scholar]
  • 435.Huang JF, Zhang Y, Rittenhouse KD, Pickering EH, McDowell MT. Evaluations of tear protein markers in dry eye disease: repeatability of measurement and correlation with disease. Invest Ophthalmol Vis Sci. 2012;53:4556–64. doi: 10.1167/iovs.11-9054. [DOI] [PubMed] [Google Scholar]
  • 436.Yoon KC, Park CS, You IC, Choi HJ, Lee KH, Im SK, et al. Expression of CXCL9, -10, -11, and CXCR3 in the tear film and ocular surface of patients with dry eye syndrome. Invest Ophthalmol Vis Sci. 2010;51:643–50. doi: 10.1167/iovs.09-3425. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 437.Sack R, Conradi L, Beaton A, Sathe S, McNamara N, Leonardi A. Antibody array characterization of inflammatory mediators in allergic and normal tears in the open and closed eye environments. Exp Eye Res. 2007;85:528–38. doi: 10.1016/j.exer.2007.07.004. [DOI] [PubMed] [Google Scholar]
  • 438.Sanghi S, Kumar R, Lumsden A, Dickinson D, Klepeis V, Trinkaus-Randall V, et al. cDNA and genomic cloning of lacritin, a novel secretion enhancing factor from the human lacrimal gland. J Mol Biol. 2001;310:127–39. doi: 10.1006/jmbi.2001.4748. [DOI] [PubMed] [Google Scholar]
  • 439.LaFrance MW, Kehinde LE, Fullard RJ. Multiple cytokine analysis in human tears: an optimized procedure for cytometric bead-based assay. Curr Eye Res. 2008;33:525–44. doi: 10.1080/02713680802190085. [DOI] [PubMed] [Google Scholar]
  • 440.Willcox MD, Zhao Z, Naduvilath T, Lazon de la Jara P. Cytokine changes in tears and relationship to contact lens discomfort. Mol Vis. 2015;21:293–305. [PMC free article] [PubMed] [Google Scholar]
  • 441.Ananthi S, Venkatesh Prajna N, Lalitha P, Valarnila M, Dharmalingam K. Pathogen induced changes in the protein profile of human tears from Fusarium keratitis patients. PLoS One. 2013;8:e53018. doi: 10.1371/journal.pone.0053018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 442.Grus FH, Podust VN, Bruns K, Lackner K, Fu S, Dalmasso EA, et al. SELDI-TOF-MS ProteinChip array profiling of tears from patients with dry eye. Invest Ophthalmol Vis Sci. 2005;46:863–76. doi: 10.1167/iovs.04-0448. [DOI] [PubMed] [Google Scholar]
  • 443.Koo BS, Lee DY, Ha HS, Kim JC, Kim CW. Comparative analysis of the tear protein expression in blepharitis patients using two-dimensional electrophoresis. J Proteom Res. 2005;4:719–24. doi: 10.1021/pr0498133. [DOI] [PubMed] [Google Scholar]
  • 444.Lee SH, Park MY, Kim KW, Wee SW, Kim JC. Zinc finger protein in severe dry eye syndrome. Curr Eye Res. 2014;39:431–8. doi: 10.3109/02713683.2013.851705. [DOI] [PubMed] [Google Scholar]
  • 445.Lei Z, Beuerman RW, Chew AP, Koh SK, Cafaro TA, Urrets-Zavalia EA, et al. Quantitative analysis of N-linked glycoproteins in tear fluid of climatic droplet keratopathy by glycopeptide capture and iTRAQ. J Proteom Res. 2009;8:1992–2003. doi: 10.1021/pr800962q. [DOI] [PubMed] [Google Scholar]
  • 446.Matheis N, Grus FH, Breitenfeld M, Knych I, Funke S, Pitz S, et al. Proteomics differentiate between thyroid-associated orbitopathy and dry eye syndrome. Invest Ophthalmol Vis Sci. 2015;56:2649–56. doi: 10.1167/iovs.15-16699. [DOI] [PubMed] [Google Scholar]
  • 447.Srinivasan S, Thangavelu M, Zhang L, Green KB, Nichols KK. iTRAQ quantitative proteomics in the analysis of tears in dry eye patients. Invest Ophthalmol Vis Sci. 2012;53:5052–9. doi: 10.1167/iovs.11-9022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 448.Tong L, Zhou L, Beuerman RW, Zhao SZ, Li XR. Association of tear proteins with Meibomian gland disease and dry eye symptoms. Br J Ophthalmol. 2011;95:848–52. doi: 10.1136/bjo.2010.185256. [DOI] [PubMed] [Google Scholar]
  • 449.Guyette N, Williams L, Tran MT, Than T, Bradley J, Kehinde L, et al. Comparison of low-abundance biomarker levels in capillary-collected non-stimulated tears and washout tears of aqueous-deficient and normal patients. Invest Ophthalmol Vis Sci. 2013;54:3729–37. doi: 10.1167/iovs.12-11431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 450.Versura P, Bavelloni A, Grillini M, Fresina M, Campos EC. Diagnostic performance of a tear protein panel in early dry eye. Mol Vis. 2013;19:1247–57. [PMC free article] [PubMed] [Google Scholar]
  • 451.Versura P, Nanni P, Bavelloni A, Blalock WL, Piazzi M, Roda A, et al. Tear proteomics in evaporative dry eye disease. Eye (Lond) 2010;24:1396–402. doi: 10.1038/eye.2010.7. [DOI] [PubMed] [Google Scholar]
  • 452.McNamara NA, Ge S, Lee SM, Enghauser AM, Kuehl L, Chen FY, et al. Reduced Levels of Tear Lacritin Are Associated With Corneal Neuropathy in Patients With the Ocular Component of Sjogren’s Syndrome. Invest Ophthalmol Vis Sci. 2016;57:5237–43. doi: 10.1167/iovs.16-19309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 453.Masoudi S, Stapleton FJ, Willcox MD. Contact Lens-Induced Discomfort and Protein Changes in Tears. Optom Vis Sci. 2016;93:955–62. doi: 10.1097/OPX.0000000000000888. [DOI] [PubMed] [Google Scholar]
  • 454.Domingo I, Coll J, Ribas-Montobio J, Marrugat J, Rubies-Prat J. Lacrimal immunoglobulins in rheumatoid arthritis patients with or without Sjogren’s syndrome. Ophthalmologica. 1998;212:30–3. doi: 10.1159/000027255. [DOI] [PubMed] [Google Scholar]
  • 455.Hamm-Alvarez SF, Janga SR, Edman MC, Madrigal S, Shah M, Frousiakis SE, et al. Tear cathepsin S as a candidate biomarker for Sjogren’s syndrome. Arthritis Rheumatol. 2014;66:1872–81. doi: 10.1002/art.38633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 456.Posa A, Brauer L, Schicht M, Garreis F, Beileke S, Paulsen F. Schirmer strip vs. capillary tube method: non-invasive methods of obtaining proteins from tear fluid. Ann Anat. 2013;195:137–42. doi: 10.1016/j.aanat.2012.10.001. [DOI] [PubMed] [Google Scholar]
  • 457.Perumal N, Funke S, Wolters D, Pfeiffer N, Grus FH. Characterization of human reflex tear proteome reveals high expression of lacrimal proline-rich protein 4 (PRR4) Proteomics. 2015;15:3370–81. doi: 10.1002/pmic.201400239. [DOI] [PubMed] [Google Scholar]
  • 458.Fullard RJ, Snyder C. Protein levels in nonstimulated and stimulated tears of normal human subjects. Invest Ophthalmol Vis Sci. 1990;31:1119–26. [PubMed] [Google Scholar]
  • 459.Fullard RJ, Tucker DL. Changes in human tear protein levels with progressively increasing stimulus. Invest Ophthalmol Vis Sci. 1991;32:2290–301. [PubMed] [Google Scholar]
  • 460.Willcox MD, Morris CA, Thakur A, Sack RA, Wickson J, Boey W. Complement and complement regulatory proteins in human tears. Invest Ophthalmol Vis Sci. 1997;38:1–8. [PubMed] [Google Scholar]
  • 461.Masoudi S, Zhao Z, Stapleton F, Willcox M. Contact Lens-Induced Discomfort and Inflammatory Mediator Changes in Tears. Eye Contact Lens. 2017;43:40–5. doi: 10.1097/ICL.0000000000000237. [DOI] [PubMed] [Google Scholar]
  • 462.Ferguson RE, Carroll HP, Harris A, Maher ER, Selby PJ, Banks RE. Housekeeping proteins: a preliminary study illustrating some limitations as useful references in protein expression studies. Proteomics. 2005;5:566–71. doi: 10.1002/pmic.200400941. [DOI] [PubMed] [Google Scholar]
  • 463.Opsahl JA, Vaudel M, Guldbrandsen A, Aasebo E, Van Pesch V, Franciotta D, et al. Label-free analysis of human cerebrospinal fluid addressing various normalization strategies and revealing protein groups affected by multiple sclerosis. Proteomics. 2016;16:1154–65. doi: 10.1002/pmic.201500284. [DOI] [PubMed] [Google Scholar]
  • 464.Pflugfelder SC, Jones D, Ji Z, Afonso A, Monroy D. Altered cytokine balance in the tear fluid and conjunctiva of patients with Sjogren’s syndrome kerato-conjunctivitis sicca. Curr Eye Res. 1999;19:201–11. doi: 10.1076/ceyr.19.3.201.5309. [DOI] [PubMed] [Google Scholar]
  • 465.Tishler M, Yaron I, Geyer O, Shirazi I, Naftaliev E, Yaron M. Elevated tear interleukin-6 levels in patients with Sjogren syndrome. Ophthalmology. 1998;105:2327–9. doi: 10.1016/S0161-6420(98)91236-2. [DOI] [PubMed] [Google Scholar]
  • 466.Danjo Y, Lee M, Horimoto K, Hamano T. Ocular surface damage and tear lactoferrin in dry eye syndrome. Acta Ophthalmol (Copenh) 1994;72:433–7. doi: 10.1111/j.1755-3768.1994.tb02791.x. [DOI] [PubMed] [Google Scholar]
  • 467.Navone R, Lunardi C, Gerli R, Tinazzi E, Peterlana D, Bason C, et al. Identification of tear lipocalin as a novel autoantigen target in Sjogren’s syndrome. J Autoimmun. 2005;25:229–34. doi: 10.1016/j.jaut.2005.09.021. [DOI] [PubMed] [Google Scholar]
  • 468.Yamada M, Mochizuki H, Kawai M, Tsubota K, Bryce TJ. Decreased tear lipocalin concentration in patients with meibomian gland dysfunction. Br J Ophthalmol. 2005;89:803–5. doi: 10.1136/bjo.2004.055822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 469.Zhao H, Jumblatt JE, Wood TO, Jumblatt MM. Quantification of MUC5AC protein in human tears. Cornea. 2001;20:873–7. doi: 10.1097/00003226-200111000-00019. [DOI] [PubMed] [Google Scholar]
  • 470.Aho VV, Nevalainen TJ, Saari KM. Group IIA phospholipase A2 content of tears in patients with keratoconjunctivitis sicca. Graefes Arch Clin Exp Ophthalmol. 2002;240:521–3. doi: 10.1007/s00417-002-0477-8. [DOI] [PubMed] [Google Scholar]
  • 471.Chen D, Wei Y, Li X, Epstein S, Wolosin JM, Asbell P. sPLA2-IIa is an inflammatory mediator when the ocular surface is compromised. Exp Eye Res. 2009;88:880–8. doi: 10.1016/j.exer.2008.11.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 472.Acera A, Suarez T, Rodriguez-Agirretxe I, Vecino E, Duran JA. Changes in tear protein profile in patients with conjunctivochalasis. Cornea. 2011;30:42–9. doi: 10.1097/ICO.0b013e3181dea7d7. [DOI] [PubMed] [Google Scholar]
  • 473.Aho VV, Paavilainen V, Nevalainen TJ, Peuravuori H, Saari KM. Diurnal variation in group IIa phospholipase A2 content in tears of contact lens wearers and normal controls. Graefes Arch Clin Exp Ophthalmol. 2003;241:85–8. doi: 10.1007/s00417-002-0607-3. [DOI] [PubMed] [Google Scholar]
  • 474.Maria NI, Vogelsang P, Versnel MA. The clinical relevance of animal models in Sjogren’s syndrome: the interferon signature from mouse to man. Arthritis Res Ther. 2015;17:172. doi: 10.1186/s13075-015-0678-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 475.Bian F, Barbosa FL, Corrales RM, Pelegrino FS, Volpe EA, Pflugfelder SC, et al. Altered balance of interleukin-13/interferon-gamma contributes to lacrimal gland destruction and secretory dysfunction in CD25 knockout model of Sjogren’s syndrome. Arthritis Res Ther. 2015;17:53. doi: 10.1186/s13075-015-0582-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 476.Lee SY, Han SJ, Nam SM, Yoon SC, Ahn JM, Kim TI, et al. Analysis of tear cytokines and clinical correlations in Sjogren syndrome dry eye patients and non-Sjogren syndrome dry eye patients. Am J Ophthalmol. 2013;156:247–53. doi: 10.1016/j.ajo.2013.04.003. [DOI] [PubMed] [Google Scholar]
  • 477.Afonso AA, Sobrin L, Monroy DC, Selzer M, Lokeshwar B, Pflugfelder SC. Tear fluid gelatinase B activity correlates with IL-1alpha concentration and fluorescein clearance in ocular rosacea. Invest Ophthalmol Vis Sci. 1999;40:2506–12. [PubMed] [Google Scholar]
  • 478.Jones L, Downie LE, Korb D, Benitez-del-Castillo JM, Dana R, Deng SX, et al. TFOS DEWS II Management and Therapy report. Ocul Surf. 2017;15:575–628. doi: 10.1016/j.jtos.2017.05.006. [DOI] [PubMed] [Google Scholar]
  • 479.Higuchi A, Ueno R, Shimmura S, Suematsu M, Dogru M, Tsubota K. Albumin rescues ocular epithelial cells from cell death in dry eye. Curr Eye Res. 2007;32:83–8. doi: 10.1080/02713680601147690. [DOI] [PubMed] [Google Scholar]
  • 480.Pattamatta U, Willcox M, Stapleton F, Garrett Q. Bovine lactoferrin promotes corneal wound healing and suppresses IL-1 expression in alkali wounded mouse cornea. Curr Eye Res. 2013;38:1110–7. doi: 10.3109/02713683.2013.811259. [DOI] [PubMed] [Google Scholar]
  • 481.Wang N, Zimmerman K, Raab RW, McKown RL, Hutnik CM, Talla V, et al. Lacritin rescues stressed epithelia via rapid forkhead box O3 (FOXO3)-associated autophagy that restores metabolism. J Biol Chem. 2013;288:18146–61. doi: 10.1074/jbc.M112.436584. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 482.Wang W, Jashnani A, Aluri SR, Gustafson JA, Hsueh PY, Yarber F, et al. A thermo-responsive protein treatment for dry eyes. J Control Release. 2015;199:156–67. doi: 10.1016/j.jconrel.2014.11.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 483.Seki JT, Sakurai N, Moldenhauer S, Dam J, Atenafu EG, Yip PM, et al. Human albumin eye drops as a therapeutic option for the management of keratoconjunctivitis sicca secondary to chronic graft-versus-host disease after stem-cell allografting. Curr Oncol. 2015;22:e357–63. doi: 10.3747/co.22.2620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 484.Ashby B, Garrett Q, Willcox M. Bovine lactoferrin structures promoting corneal epithelial wound healing in vitro. Invest Ophthalmol Vis Sci. 2011;52:2719–26. doi: 10.1167/iovs.10-6352. [DOI] [PubMed] [Google Scholar]
  • 485.Pastori V, Tavazzi S, Lecchi M. Lactoferrin-loaded contact lenses: eye protection against oxidative stress. Cornea. 2015;34:693–7. doi: 10.1097/ICO.0000000000000435. [DOI] [PubMed] [Google Scholar]
  • 486.Kawashima M, Kawakita T, Inaba T, Okada N, Ito M, Shimmura S, et al. Dietary lactoferrin alleviates age-related lacrimal gland dysfunction in mice. PLoS One. 2012;7:e33148. doi: 10.1371/journal.pone.0033148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 487.Kawashima M, Nakamura S, Izuta Y, Inoue S, Tsubota K. Dietary Supplementation with a Combination of Lactoferrin, Fish Oil, and Enterococcus faecium WB2000 for Treating Dry Eye: A Rat Model and Human Clinical Study. Ocul Surf. 2016;14:255–63. doi: 10.1016/j.jtos.2015.12.005. [DOI] [PubMed] [Google Scholar]
  • 488.Devendra J, Singh S. Effect of oral lactoferrin on cataract surgery induced dry eye: a randomised controlled trial. J Clin Diagn Res. 2015;9:NC06–9. doi: 10.7860/JCDR/2015/15797.6670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 489.Vijmasi T, Chen FY, Balasubbu S, Gallup M, McKown RL, Laurie GW, et al. Topical administration of lacritin is a novel therapy for aqueous-deficient dry eye disease. Invest Ophthalmol Vis Sci. 2014;55:5401–9. doi: 10.1167/iovs.14-13924. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 490.Feng MM, Baryla J, Liu H, Laurie GW, McKown RL, Ashki N, et al. Cytopro-tective effect of lacritin on human corneal epithelial cells exposed to benzalkonium chloride in vitro. Curr Eye Res. 2014;39:604–10. doi: 10.3109/02713683.2013.859275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 491.Samudre S, Lattanzio FA, Jr, Lossen V, Hosseini A, Sheppard JD, Jr, McKown RL, et al. Lacritin, a novel human tear glycoprotein, promotes sustained basal tearing and is well tolerated. Invest Ophthalmol Vis Sci. 2011;52:6265–70. doi: 10.1167/iovs.10-6220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 492.Wang Z, Chen Z, Yang Q, Jiang Y, Lin L, Liu X, et al. Vitelline membrane outer layer 1 homolog interacts with lysozyme C and promotes the stabilization of tear film. Invest Ophthalmol Vis Sci. 2014;55:6722–7. doi: 10.1167/iovs.14-14491. [DOI] [PubMed] [Google Scholar]
  • 493.Sambursky R, Davitt WF, 3rd, Latkany R, Tauber S, Starr C, Friedberg M, et al. Sensitivity and specificity of a point-of-care matrix metalloproteinase 9 immunoassay for diagnosing inflammation related to dry eye. JAMA Ophthalmol. 2013;131:24–8. doi: 10.1001/jamaophthalmol.2013.561. [DOI] [PubMed] [Google Scholar]
  • 494.D’Souza S, Petznick A, Tong L, Hall RC, Rosman M, Chan C, et al. Comparative analysis of two femtosecond LASIK platforms using iTRAQ quantitative proteomics. Invest Ophthalmol Vis Sci. 2014;55:3396–402. doi: 10.1167/iovs.14-14113. [DOI] [PubMed] [Google Scholar]
  • 495.Nättinen J, Jylhä A, Aapola U, Beuerman RW, Nykter M, Kesseli J, et al. Effects of preserved and preservative free glaucoma drugs on proteomic expression levels in corneal and conjunctival epithelial cells in vitro. Invest Ophthalmol Vis Sci. 2015;56 E-Abstract 221. [Google Scholar]
  • 496.Zimmermann N, Erb C. Immunoassay for matrix metalloproteinase-9 in the tear film of patients with pseudoexfoliation syndrome - a pilot study. Klin Monbl Augenheilkd. 2013;230:804–7. doi: 10.1055/s-0032-1328410. [DOI] [PubMed] [Google Scholar]
  • 497.Sambursky R, Davitt WF, 3rd, Friedberg M, Tauber S. Prospective, multicenter, clinical evaluation of point-of-care matrix metalloproteinase-9 test for confirming dry eye disease. Cornea. 2014;33:812–8. doi: 10.1097/ICO.0000000000000175. [DOI] [PubMed] [Google Scholar]
  • 498.Schargus M, Ivanova S, Kakkassery V, Dick HB, Joachim S. Correlation of Tear Film Osmolarity and 2 Different MMP-9 Tests With Common Dry Eye Tests in a Cohort of Non-Dry Eye Patients. Cornea. 2015;34:739–44. doi: 10.1097/ICO.0000000000000449. [DOI] [PubMed] [Google Scholar]
  • 499.Lanza NL, McClellan AL, Batawi H, Felix ER, Sarantopoulos KD, Levitt RC, et al. Dry Eye Profiles in Patients with a Positive Elevated Surface Matrix Metalloproteinase 9 Point-of-Care Test Versus Negative Patients. Ocul Surf. 2016;14:216–23. doi: 10.1016/j.jtos.2015.12.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 500.Zhou L, Beuerman RW, Foo Y, Liu S, Ang LP, Tan DT. Characterisation of human tear proteins using high-resolution mass spectrometry. Ann Acad Med Singapore. 2006;35:400–7. [PubMed] [Google Scholar]
  • 501.Wei Y, Gadaria-Rathod N, Epstein S, Asbell P. Tear cytokine profile as a noninvasive biomarker of inflammation for ocular surface diseases: standard operating procedures. Invest Ophthalmol Vis Sci. 2013;54:8327–36. doi: 10.1167/iovs.13-12132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 502.Laguna MF, Sanza FJ, Soria J, Jara M, Lavin A, Casquel R, et al. Label-free biosensing by means of BICELLs for dry eye. Sensors Actuators B. 2014;203:209–12. [Google Scholar]
  • 503.Laguna MF, Holgado M, Hernandez AL, Santamaria B, Lavin A, Soria J, et al. Antigen-antibody affinity for dry eye biomarkers by label free biosensing. Comparison with the ELISA technique. Sensors (Basel) 2015;15:19819–29. doi: 10.3390/s150819819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 504.Perumal N, Funke S, Pfeiffer N, Grus FH. Characterization of lacrimal proline-rich protein 4 (PRR4) in human tear proteome. Proteomics. 2014;14:1698–709. doi: 10.1002/pmic.201300039. [DOI] [PubMed] [Google Scholar]
  • 505.Aragona P, Aguennouz M, Rania L, Postorino E, Sommario MS, Roszkowska AM, et al. Matrix metalloproteinase 9 and transglutaminase 2 expression at the ocular surface in patients with different forms of dry eye disease. Ophthalmology. 2015;122:62–71. doi: 10.1016/j.ophtha.2014.07.048. [DOI] [PubMed] [Google Scholar]
  • 506.Csosz E, Mesko B, Fesus L. Transdab wiki: the interactive transglutaminase substrate database on web 2. 0 surface. Amino Acids. 2009;36:615–7. doi: 10.1007/s00726-008-0121-y. [DOI] [PubMed] [Google Scholar]
  • 507.Velez VF, Romano JA, McKown RL, Green K, Zhang L, Raab RW, et al. Tissue transglutaminase is a negative regulator of monomeric lacritin bioactivity. Invest Ophthalmol Vis Sci. 2013;54:2123–32. doi: 10.1167/iovs.12-11488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 508.Azharuddin M, Khandelwal J, Datta H, Dasgupta AK, Raja SO. Dry eye: a protein conformational disease. Invest Ophthalmol Vis Sci. 2015;56:1423–9. doi: 10.1167/iovs.14-15992. [DOI] [PubMed] [Google Scholar]
  • 509.Faheem S, Kim SH, Nguyen J, Neravetla S, Ball M, Foulks GN, et al. Wax-tear and meibum protein, wax-beta-carotene interactions in vitro using infrared spectroscopy. Exp Eye Res. 2012;100:32–9. doi: 10.1016/j.exer.2012.04.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 510.Setala NL, Holopainen JM, Metso J, Yohannes G, Hiidenhovi J, Andersson LC, et al. Interaction of phospholipid transfer protein with human tear fluid mucins. J Lipid Res. 2010;51:3126–34. doi: 10.1194/jlr.M006239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 511.Mauris J, Mantelli F, Woodward AM, Cao Z, Bertozzi CR, Panjwani N, et al. Modulation of ocular surface glycocalyx barrier function by a galectin-3 N-terminal deletion mutant and membrane-anchored synthetic glycopolymers. PLoS One. 2013;8:e72304. doi: 10.1371/journal.pone.0072304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 512.Jeong S, Ledee DR, Gordon GM, Itakura T, Patel N, Martin A, et al. Interaction of clusterin and matrix metalloproteinase-9 and its implication for epithelial homeostasis and inflammation. Am J Pathol. 2012;180:2028–39. doi: 10.1016/j.ajpath.2012.01.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 513.Bauskar A, Mack WJ, Mauris J, Argueso P, Heur M, Nagel BA, et al. Clusterin seals the ocular surface barrier in mouse dry eye. PLoS One. 2015;10:e0138958. doi: 10.1371/journal.pone.0138958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 514.Butovich IA, Millar TJ, Ham BM. Understanding and analyzing meibomian lipids–a review. Curr Eye Res. 2008;33:405–20. doi: 10.1080/02713680802018419. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 515.Gasymov OK, Abduragimov AR, Yusifov TN, Glasgow BJ. Binding studies of tear lipocalin: the role of the conserved tryptophan in maintaining structure, stability and ligand affinity. Biochim Biophys Acta. 1999;1433:307–20. doi: 10.1016/s0167-4838(99)00133-8. [DOI] [PubMed] [Google Scholar]
  • 516.Glasgow BJ, Abduragimov AR, Farahbakhsh ZT, Faull KF, Hubbell WL. Tear lipocalins bind a broad array of lipid ligands. Curr Eye Res. 1995;14:363–72. doi: 10.3109/02713689508999934. [DOI] [PubMed] [Google Scholar]
  • 517.Glasgow BJ, Marshall G, Gasymov OK, Abduragimov AR, Yusifov TN, Knobler CM. Tear lipocalins: potential lipid scavengers for the corneal surface. Invest Ophthalmol Vis Sci. 1999;40:3100–7. [PubMed] [Google Scholar]
  • 518.Glasson M, Stapleton F, Willcox M. Lipid, lipase and lipocalin differences between tolerant and intolerant contact lens wearers. Curr Eye Res. 2002;25:227–35. doi: 10.1076/ceyr.25.4.227.13482. [DOI] [PubMed] [Google Scholar]
  • 519.Caffery B, Joyce E, Boone A, Slomovic A, Simpson T, Jones L, et al. Tear lipocalin and lysozyme in Sjogren and non-Sjogren dry eye. Optom Vis Sci. 2008;85:661–7. doi: 10.1097/OPX.0b013e318181ae4f. [DOI] [PubMed] [Google Scholar]
  • 520.Brauer L, Johl M, Borgermann J, Pleyer U, Tsokos M, Paulsen FP. Detection and localization of the hydrophobic surfactant proteins B and C in human tear fluid and the human lacrimal system. Curr Eye Res. 2007;32:931–8. doi: 10.1080/02713680701694369. [DOI] [PubMed] [Google Scholar]
  • 521.Holzfeind P, Merschak P, Dieplinger H, Redl B. The human lacrimal gland synthesizes apolipoprotein D mRNA in addition to tear prealbumin mRNA, both species encoding members of the lipocalin superfamily. Exp Eye Res. 1995;61:495–500. doi: 10.1016/s0014-4835(05)80145-9. [DOI] [PubMed] [Google Scholar]
  • 522.Molloy MP, Bolis S, Herbert BR, Ou K, Tyler MI, van Dyk DD, et al. Establishment of the human reflex tear two-dimensional polyacrylamide gel electrophoresis reference map: new proteins of potential diagnostic value. Electrophoresis. 1997;18:2811–5. doi: 10.1002/elps.1150181516. [DOI] [PubMed] [Google Scholar]
  • 523.Remington SG, Crow JM, Nelson JD. Secretoglobins: lacrimal gland-specific rabbit lipophilin mRNAs. Invest Ophthalmol Vis Sci. 2008;49:2856–62. doi: 10.1167/iovs.07-1496. [DOI] [PubMed] [Google Scholar]
  • 524.Lehrer RI, Xu G, Abduragimov A, Dinh NN, Qu XD, Martin D, et al. Lipophilin, a novel heterodimeric protein of human tears. FEBS Lett. 1998;432:163–7. doi: 10.1016/s0014-5793(98)00852-7. [DOI] [PubMed] [Google Scholar]
  • 525.Jauhiainen M, Setala NL, Ehnholm C, Metso J, Tervo TM, Eriksson O, et al. Phospholipid transfer protein is present in human tear fluid. Biochemistry. 2005;44:8111–6. doi: 10.1021/bi050151k. [DOI] [PubMed] [Google Scholar]
  • 526.Ma P, Beck SL, Raab RW, McKown RL, Coffman GL, Utani A, et al. Heparanase deglycanation of syndecan-1 is required for binding of the epithelial-restricted prosecretory mitogen lacritin. J Cell Biol. 2006;174:1097–106. doi: 10.1083/jcb.200511134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 527.Zhang Y, Wang N, Raab RW, McKown RL, Irwin JA, Kwon I, et al. Targeting of heparanase-modified syndecan-1 by prosecretory mitogen lacritin requires conserved core GAGAL plus heparan and chondroitin sulfate as a novel hybrid binding site that enhances selectivity. J Biol Chem. 2013;288:12090–101. doi: 10.1074/jbc.M112.422717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 528.Medina S, Dominguez-Perles R, Gil JI, Ferreres F, Gil-Izquierdo A. Metabolomics and the diagnosis of human diseases–a guide to the markers and pathophysiological pathways affected. Curr Med Chem. 2014;21:823–48. doi: 10.2174/0929867320666131119124056. [DOI] [PubMed] [Google Scholar]
  • 529.Munoz-Hernandez AM, Galbis-Estrada C, Santos-Bueso E, Cuina-Sardina R, Diaz-Valle D, Gegundez-Fernandez JA, et al. Human tear metabolome. Arch Soc Esp Oftalmol. 2016;91:157–9. doi: 10.1016/j.oftal.2015.09.010. [DOI] [PubMed] [Google Scholar]
  • 530.Galbis-Estrada C, Martinez-Castillo S, Morales JM, Vivar-Llopis B, Monleon D, Diaz-Llopis M, et al. Differential effects of dry eye disorders on metabolomic profile by 1H nuclear magnetic resonance spectroscopy. Biomed Res Int. 2014;2014:542549. doi: 10.1155/2014/542549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 531.Galbis-Estrada C, Pinazo-Duran MD, Martinez-Castillo S, Morales JM, Monleon D, Zanon-Moreno V. A metabolomic approach to dry eye disorders. The role of oral supplements with antioxidants and omega 3 fatty acids. Mol Vis. 2015;21:555–67. [PMC free article] [PubMed] [Google Scholar]
  • 532.Lindon JC, Holmes E, Bollard ME, Stanley EG, Nicholson JK. Metabonomics technologies and their applications in physiological monitoring, drug safety assessment and disease diagnosis. Biomarkers. 2004;9:1–31. doi: 10.1080/13547500410001668379. [DOI] [PubMed] [Google Scholar]
  • 533.Nicholson JK, Lindon JC, Holmes E. ‘Metabonomics’: understanding the metabolic responses of living systems to pathophysiological stimuli via multivariate statistical analysis of biological NMR spectroscopic data. Xenobiotica. 1999;29:1181–9. doi: 10.1080/004982599238047. [DOI] [PubMed] [Google Scholar]
  • 534.Chen L, Zhou L, Chan EC, Neo J, Beuerman RW. Characterization of the human tear metabolome by LC-MS/MS. J Proteom Res. 2011;10:4876–82. doi: 10.1021/pr2004874. [DOI] [PubMed] [Google Scholar]
  • 535.Want EJ, Wilson ID, Gika H, Theodoridis G, Plumb RS, Shockcor J, et al. Global metabolic profiling procedures for urine using UPLC-MS. Nat Protoc. 2010;5:1005–18. doi: 10.1038/nprot.2010.50. [DOI] [PubMed] [Google Scholar]
  • 536.Nicholson JK, Lindon JC. Systems biology: Metabonomics. Nature. 2008;455:1054–6. doi: 10.1038/4551054a. [DOI] [PubMed] [Google Scholar]
  • 537.Nakatsukasa M, Sotozono C, Shimbo K, Ono N, Miyano H, Okano A, et al. Amino Acid profiles in human tear fluids analyzed by high-performance liquid chromatography and electrospray ionization tandem mass spectrometry. Am J Ophthalmol. 2011;151:799–808. doi: 10.1016/j.ajo.2010.11.003. [DOI] [PubMed] [Google Scholar]
  • 538.Choy CK, Cho P, Chung WY, Benzie IF. Water-soluble antioxidants in human tears: effect of the collection method. Invest Ophthalmol Vis Sci. 2001;42:3130–4. [PubMed] [Google Scholar]
  • 539.Gogia R, Richer SP, Rose RC. Tear fluid content of electrochemically active components including water soluble antioxidants. Curr Eye Res. 1998;17:257–63. doi: 10.1076/ceyr.17.3.257.5213. [DOI] [PubMed] [Google Scholar]
  • 540.Baca JT, Taormina CR, Feingold E, Finegold DN, Grabowski JJ, Asher SA. Mass spectral determination of fasting tear glucose concentrations in nondiabetic volunteers. Clin Chem. 2007;53:1370–2. doi: 10.1373/clinchem.2006.078543. [DOI] [PubMed] [Google Scholar]
  • 541.Taormina CR, Baca JT, Asher SA, Grabowski JJ, Finegold DN. Analysis of tear glucose concentration with electrospray ionization mass spectrometry. J Am Soc Mass Spectrom. 2007;18:332–6. doi: 10.1016/j.jasms.2006.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 542.Pescosolido N, Imperatrice B, Koverech A, Messano M. L-carnitine and short chain ester in tears from patients with dry eye. Optom Vis Sci. 2009;86:E132–8. doi: 10.1097/OPX.0b013e318194e767. [DOI] [PubMed] [Google Scholar]
  • 543.Trope GE, Rumley AG. Catecholamine concentrations in tears. Exp Eye Res. 1984;39:247–50. doi: 10.1016/0014-4835(84)90012-5. [DOI] [PubMed] [Google Scholar]
  • 544.Pintor J, Carracedo G, Alonso MC, Bautista A, Peral A. Presence of diadenosine polyphosphates in human tears. Pflugers Arch. 2002;443:432–6. doi: 10.1007/s004240100696. [DOI] [PubMed] [Google Scholar]
  • 545.Speek AJ, van Agtmaal EJ, Saowakontha S, Schreurs WH, van Haeringen NJ. Fluorometric determination of retinol in human tear fluid using high-performance liquid chromatography. Curr Eye Res. 1986;5:841–5. doi: 10.3109/02713688609029235. [DOI] [PubMed] [Google Scholar]
  • 546.Mendelsohn ME, Abramson DH, Senft S, Servodidio CA, Gamache PH. Uric acid in the aqueous humor and tears of retinoblastoma patients. J AAPOS. 1998;2:369–71. doi: 10.1016/s1091-8531(98)90037-4. [DOI] [PubMed] [Google Scholar]
  • 547.Peral A, Carracedo G, Acosta MC, Gallar J, Pintor J. Increased levels of dia-denosine polyphosphates in dry eye. Invest Ophthalmol Vis Sci. 2006;47:4053–8. doi: 10.1167/iovs.05-0980. [DOI] [PubMed] [Google Scholar]
  • 548.Carracedo G, Peral A, Pintor J. Diadenosine polyphosphates in tears of Sjogren syndrome patients. Invest Ophthalmol Vis Sci. 2010;51:5452–9. doi: 10.1167/iovs.09-5088. [DOI] [PubMed] [Google Scholar]
  • 549.Chhadva P, Lee T, Sarantopoulos CD, Hackam AS, McClellan AL, Felix ER, et al. Human tear serotonin levels correlate with symptoms and signs of dry eye. Ophthalmology. 2015;122:1675–80. doi: 10.1016/j.ophtha.2015.04.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 550.Barabino S, Dana MR. Animal models of dry eye: a critical assessment of opportunities and limitations. Invest Ophthalmol Vis Sci. 2004;45:1641–6. doi: 10.1167/iovs.03-1055. [DOI] [PubMed] [Google Scholar]
  • 551.Walcott B, Matthews G, Brink P. Differences in stimulus induced calcium increases in lacrimal gland acinar cells from normal and NZB/NZW F1 female mice. Curr Eye Res. 2002;25:253–60. doi: 10.1076/ceyr.25.4.253.13489. [DOI] [PubMed] [Google Scholar]
  • 552.Ding C, MacVeigh M, Pidgeon M, da Costa SR, Wu K, Hamm-Alvarez SF, et al. Unique ultrastructure of exorbital lacrimal glands in male NOD and BALB/c mice. Curr Eye Res. 2006;31:13–22. doi: 10.1080/02713680500428613. [DOI] [PubMed] [Google Scholar]
  • 553.You IC, Bian F, Volpe EA, de Paiva CS, Pflugfelder SC. Age-Related Conjunctival Disease in the C57BL/6. NOD-Aec1Aec2 Mouse Model of Sjogren Syndrome Develops Independent of Lacrimal Dysfunction. Invest Ophthalmol Vis Sci. 2015;56:2224–33. doi: 10.1167/iovs.14-15668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 554.De Paiva CS, Hwang CS, Pitcher JD, 3rd, Pangelinan SB, Rahimy E, Chen W, et al. Age-related T-cell cytokine profile parallels corneal disease severity in Sjogren’s syndrome-like keratoconjunctivitis sicca in CD25KO mice. Rheumatology (Oxford) 2010;49:246–58. doi: 10.1093/rheumatology/kep357. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 555.Dursun D, Wang M, Monroy D, Li DQ, Lokeshwar BL, Stern ME, et al. A mouse model of keratoconjunctivitis sicca. Invest Ophthalmol Vis Sci. 2002;43:632–8. [PubMed] [Google Scholar]
  • 556.Barabino S, Shen L, Chen L, Rashid S, Rolando M, Dana MR. The controlled-environment chamber: a new mouse model of dry eye. Invest Ophthalmol Vis Sci. 2005;46:2766–71. doi: 10.1167/iovs.04-1326. [DOI] [PubMed] [Google Scholar]
  • 557.Coursey TG, Henriksson JT, Marcano DC, Shin CS, Isenhart LC, Ahmed F, et al. Dexamethasone nanowafer as an effective therapy for dry eye disease. J Control Release. 2015;213:168–74. doi: 10.1016/j.jconrel.2015.07.007. [DOI] [PubMed] [Google Scholar]
  • 558.Coursey TG, Bohat R, Barbosa FL, Pflugfelder SC, de Paiva CS. Desiccating stress-induced chemokine expression in the epithelium is dependent on upregulation of NKG2D/RAE-1 and release of IFN-gamma in experimental dry eye. J Immunol. 2014;193:5264–72. doi: 10.4049/jimmunol.1400016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 559.Suhalim JL, Parfitt GJ, Xie Y, De Paiva CS, Pflugfelder SC, Shah TN, et al. Effect of desiccating stress on mouse meibomian gland function. Ocul Surf. 2014;12:59–68. doi: 10.1016/j.jtos.2013.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 560.Zhang X, De Paiva CS, Su Z, Volpe EA, Li DQ, Pflugfelder SC. Topical inter-feron-gamma neutralization prevents conjunctival goblet cell loss in experimental murine dry eye. Exp Eye Res. 2014;118:117–24. doi: 10.1016/j.exer.2013.11.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 561.Zhang X, Schaumburg CS, Coursey TG, Siemasko KF, Volpe EA, Gandhi NB, et al. CD8(+) cells regulate the T helper-17 response in an experimental murine model of Sjogren syndrome. Mucosal Immunol. 2014;7:417–27. doi: 10.1038/mi.2013.61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 562.Henriksson JT, De Paiva CS, Farley W, Pflugfelder SC, Burns AR, Bergmanson JP. Morphologic alterations of the palpebral conjunctival epithelium in a dry eye model. Cornea. 2013;32:483–90. doi: 10.1097/ICO.0b013e318265682c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 563.de Paiva CS, Schwartz CE, Gjorstrup P, Pflugfelder SC. Resolvin E1 (RX-10001) reduces corneal epithelial barrier disruption and protects against goblet cell loss in a murine model of dry eye. Cornea. 2012;31:1299–303. doi: 10.1097/ICO.0b013e31823f789e. [DOI] [PubMed] [Google Scholar]
  • 564.De Paiva CS, Volpe EA, Gandhi NB, Zhang X, Zheng X, Pitcher JD, 3rd, et al. Disruption of TGF-beta signaling improves ocular surface epithelial disease in experimental autoimmune keratoconjunctivitis sicca. PLoS One. 2011;6:e29017. doi: 10.1371/journal.pone.0029017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 565.Chang E, McClellan AJ, Farley WJ, Li DQ, Pflugfelder SC, De Paiva CS. Biodegradable PLGA-Based Drug Delivery Systems for Modulating Ocular Surface Disease under Experimental Murine Dry Eye. J Clin Exp Ophthalmol. 2011;2:1–6. doi: 10.4172/2155-9570.1000191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 566.Schaumburg CS, Siemasko KF, De Paiva CS, Wheeler LA, Niederkorn JY, Pflugfelder SC, et al. Ocular surface APCs are necessary for autoreactive T cell-mediated experimental autoimmune lacrimal keratoconjunctivitis. J Immunol. 2011;187:3653–62. doi: 10.4049/jimmunol.1101442. [DOI] [PubMed] [Google Scholar]
  • 567.Zhang X, Chen W, De Paiva CS, Volpe EA, Gandhi NB, Farley WJ, et al. Desiccating stress induces CD4+ T-cell-mediated Sjogren’s syndrome-like corneal epithelial apoptosis via activation of the extrinsic apoptotic pathway by interferon-gamma. Am J Pathol. 2011;179:1807–14. doi: 10.1016/j.ajpath.2011.06.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 568.De Paiva CS, Chotikavanich S, Pangelinan SB, Pitcher JD, 3rd, Fang B, Zheng X, et al. IL-17 disrupts corneal barrier following desiccating stress. Mucosal Immunol. 2009;2:243–53. doi: 10.1038/mi.2009.5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 569.Siemasko KF, Gao J, Calder VL, Hanna R, Calonge M, Pflugfelder SC, et al. In vitro expanded CD4+CD25+Foxp3+ regulatory T cells maintain a normal phenotype and suppress immune-mediated ocular surface inflammation. Invest Ophthalmol Vis Sci. 2008;49:5434–40. doi: 10.1167/iovs.08-2075. [DOI] [PubMed] [Google Scholar]
  • 570.Stewart P, Chen Z, Farley W, Olmos L, Pflugfelder SC. Effect of experimental dry eye on tear sodium concentration in the mouse. Eye Contact Lens. 2005;31:175–8. doi: 10.1097/01.icl.0000161705.19602.c9. [DOI] [PubMed] [Google Scholar]
  • 571.Strong B, Farley W, Stern ME, Pflugfelder SC. Topical cyclosporine inhibits conjunctival epithelial apoptosis in experimental murine keratoconjunctivitis sicca. Cornea. 2005;24:80–5. doi: 10.1097/01.ico.0000133994.22392.47. [DOI] [PubMed] [Google Scholar]
  • 572.Li Z, Woo JM, Chung SW, Kwon MY, Choi JS, Oh HJ, et al. Therapeutic effect of topical adiponectin in a mouse model of desiccating stress-induced dry eye. Invest Ophthalmol Vis Sci. 2013;54:155–62. doi: 10.1167/iovs.12-10648. [DOI] [PubMed] [Google Scholar]
  • 573.Corrales RM, Stern ME, De Paiva CS, Welch J, Li DQ, Pflugfelder SC. Desiccating stress stimulates expression of matrix metalloproteinases by the corneal epithelium. Invest Ophthalmol Vis Sci. 2006;47:3293–302. doi: 10.1167/iovs.05-1382. [DOI] [PubMed] [Google Scholar]
  • 574.Corrales RM, Villarreal A, Farley W, Stern ME, Li DQ, Pflugfelder SC. Strain-related cytokine profiles on the murine ocular surface in response to desiccating stress. Cornea. 2007;26:579–84. doi: 10.1097/ICO.0b013e318033a729. [DOI] [PubMed] [Google Scholar]
  • 575.Ozyildirim AM, Wistow GJ, Gao J, Wang J, Dickinson DP, Frierson HF, Jr, et al. The lacrimal gland transcriptome is an unusually rich source of rare and poorly characterized gene transcripts. Invest Ophthalmol Vis Sci. 2005;46:1572–80. doi: 10.1167/iovs.04-1380. [DOI] [PubMed] [Google Scholar]
  • 576.Zhou L, Wei R, Zhao P, Koh SK, Beuerman RW, Ding C. Proteomic analysis revealed the altered tear protein profile in a rabbit model of Sjogren’s syndrome-associated dry eye. Proteomics. 2013;13:2469–81. doi: 10.1002/pmic.201200230. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 577.Butovich IA, Lu H, McMahon A, Eule JC. Toward an animal model of the human tear film: biochemical comparison of the mouse, canine, rabbit, and human meibomian lipidomes. Invest Ophthalmol Vis Sci. 2012;53:6881–96. doi: 10.1167/iovs.12-10516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 578.Li N, Deng X, Gao Y, Zhang S, He M, Zhao D. Establishment of the mild, moderate and severe dry eye models using three methods in rabbits. BMC Ophthalmol. 2013;13:50. doi: 10.1186/1471-2415-13-50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 579.Bron AJ, de Paiva CS, Chauhan SK, Bonini S, Gabison EE, Jain S, et al. TFOS DEWS II Pathophysiology report. Ocul Surf. 2017;15:438–510. doi: 10.1016/j.jtos.2017.05.011. [DOI] [PubMed] [Google Scholar]
  • 580.Tan LL, Morgan P, Cai ZQ, Straughan RA. Prevalence of and risk factors for symptomatic dry eye disease in Singapore. Clin Exp Optom. 2015;98:45–53. doi: 10.1111/cxo.12210. [DOI] [PubMed] [Google Scholar]
  • 581.Paulsen AJ, Cruickshanks KJ, Fischer ME, Huang GH, Klein BE, Klein R, et al. Dry eye in the beaver dam offspring study: prevalence, risk factors, and health-related quality of life. Am J Ophthalmol. 2014;157:799–806. doi: 10.1016/j.ajo.2013.12.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 582.Yang WJ, Yang YN, Cao J, Man ZH, Yuan J, Xiao X, et al. Risk Factors for Dry Eye Syndrome: A Retrospective Case-Control Study. Optom Vis Sci. 2015;92:e199–205. doi: 10.1097/OPX.0000000000000541. [DOI] [PubMed] [Google Scholar]
  • 583.Holly FJ. Tear film physiology and contact lens wear. II. Contact lens-tear film interaction. Am J Optom Physiol Opt. 1981;58:331–41. doi: 10.1097/00006324-198104000-00012. [DOI] [PubMed] [Google Scholar]
  • 584.Nichols JJ, King-Smith PE. The impact of hydrogel lens settling on the thickness of the tears and contact lens. Invest Ophthalmol Vis Sci. 2004;45:2549–54. doi: 10.1167/iovs.04-0149. [DOI] [PubMed] [Google Scholar]
  • 585.Muselier-Mathieu A, Bron AM, Mathieu B, Souchier M, Brignole-Baudouin F, Acar N, et al. Ocular surface assessment in soft contact lens wearers; the contribution of tear osmolarity among other tests. Acta Ophthalmol. 2014;92:364–9. doi: 10.1111/aos.12103. [DOI] [PubMed] [Google Scholar]
  • 586.Rohit A, Willcox MD, Brown SH, Mitchell TW, Stapleton F. Clinical and biochemical tear lipid parameters in contact lens wearers. Optom Vis Sci. 2014;91:1384–90. doi: 10.1097/OPX.0000000000000420. [DOI] [PubMed] [Google Scholar]
  • 587.Muntz A, Subbaraman LN, Sorbara L, Jones L. Tear exchange and contact lenses: a review. J Optom. 2015;8:2–11. doi: 10.1016/j.optom.2014.12.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 588.McNamara NA, Polse KA, Bonanno JA. Fluorophotometry in contact lens research: the next step. Optom Vis Sci. 1998;75:316–22. doi: 10.1097/00006324-199805000-00020. [DOI] [PubMed] [Google Scholar]
  • 589.Paugh JR, Stapleton F, Keay L, Ho A. Tear exchange under hydrogel contact lenses: methodological considerations. Invest Ophthalmol Vis Sci. 2001;42:2813–20. [PubMed] [Google Scholar]
  • 590.Polse KA. Tear flow under hydrogel contact lenses. Invest Ophthalmol Vis Sci. 1979;18:409–13. [PubMed] [Google Scholar]
  • 591.Luo L, Li DQ, Doshi A, Farley W, Corrales RM, Pflugfelder SC. Experimental dry eye stimulates production of inflammatory cytokines and MMP-9 and activates MAPK signaling pathways on the ocular surface. Invest Ophthalmol Vis Sci. 2004;45:4293–301. doi: 10.1167/iovs.03-1145. [DOI] [PubMed] [Google Scholar]
  • 592.Heimer SR, Evans DJ, Mun JJ, Stern ME, Fleiszig SM. Surfactant protein D contributes to ocular defense against Pseudomonas aeruginosa in a murine model of dry eye disease. PLoS One. 2013;8:e65797. doi: 10.1371/journal.pone.0065797. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 593.Toshida H, Nguyen DH, Beuerman RW, Murakami A. Neurologic evaluation of acute lacrimomimetic effect of cyclosporine in an experimental rabbit dry eye model. Invest Ophthalmol Vis Sci. 2009;50:2736–41. doi: 10.1167/iovs.08-1880. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 594.Toshida H, Nguyen DH, Beuerman RW, Murakami A. Evaluation of novel dry eye model: preganglionic parasympathetic denervation in rabbit. Invest Ophthalmol Vis Sci. 2007;48:4468–75. doi: 10.1167/iovs.06-1486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 595.Wolffsohn JS, Mroczkowska S, Hunt OA, Bilkhu P, Drew T, Sheppard A. Crossover evaluation of silicone hydrogel daily disposable contact lenses. Optom Vis Sci. 2015;92:1063–8. doi: 10.1097/OPX.0000000000000706. [DOI] [PubMed] [Google Scholar]
  • 596.Bhargava R, Kumar P. Oral omega-3 fatty acid treatment for dry eye in contact lens wearers. Cornea. 2015;34:413–20. doi: 10.1097/ICO.0000000000000386. [DOI] [PubMed] [Google Scholar]
  • 597.Lopez-de la Rosa A, Martin-Montanez V, Lopez-Miguel A, Calonge M, Enriquez-de-Salamanca A, Gonzalez-Garcia MJ. Corneal Sensitivity and Inflammatory Biomarkers in Contact Lens Discomfort. Optom Vis Sci. 2016;93:892–900. doi: 10.1097/OPX.0000000000000784. [DOI] [PubMed] [Google Scholar]
  • 598.Fukui M, Yamada M, Akune Y, Shigeyasu C, Tsubota K. Fluorophotometric Analysis of the Ocular Surface Glycocalyx in Soft Contact Lens Wearers. Curr Eye Res. 2016;41:9–14. doi: 10.3109/02713683.2014.999948. [DOI] [PubMed] [Google Scholar]
  • 599.Mathers WD, Stovall D, Lane JA, Zimmerman MB, Johnson S. Menopause and tear function: the influence of prolactin and sex hormones on human tear production. Cornea. 1998;17:353–8. doi: 10.1097/00003226-199807000-00002. [DOI] [PubMed] [Google Scholar]
  • 600.Gagliano C, Caruso S, Napolitano G, Malaguarnera G, Cicinelli MV, Amato R, et al. Low levels of 17-beta-oestradiol, oestrone and testosterone correlate with severe evaporative dysfunctional tear syndrome in postmenopausal women: a case-control study. Br J Ophthalmol. 2014;98:371–6. doi: 10.1136/bjophthalmol-2012-302705. [DOI] [PubMed] [Google Scholar]
  • 601.Scuderi G, Contestabile MT, Gagliano C, Iacovello D, Scuderi L, Avitabile T. Effects of phytoestrogen supplementation in postmenopausal women with dry eye syndrome: a randomized clinical trial. Can J Ophthalmol. 2012;47:489–92. doi: 10.1016/j.jcjo.2012.08.019. [DOI] [PubMed] [Google Scholar]
  • 602.Srinivasan S, Joyce E, Boone A, Simpson T, Jones L, Senchyna M. Tear lipocalin and lysozyme concentrations in postmenopausal women. Ophthalmic Physiol Opt. 2010;30:257–66. doi: 10.1111/j.1475-1313.2010.00714.x. [DOI] [PubMed] [Google Scholar]
  • 603.Careba I, Chiva A, Totir M, Ungureanu E, Gradinaru S. Tear lipocalin, lysozyme and lactoferrin concentrations in postmenopausal women. J Med Life. 2015;8(Spec Issue):94–8. [PMC free article] [PubMed] [Google Scholar]
  • 604.Taner P, Akarsu C, Atasoy P, Bayram M, Ergin A. The effects of hormone replacement therapy on ocular surface and tear function tests in post-menopausal women. Ophthalmologica. 2004;218:257–9. doi: 10.1159/000078616. [DOI] [PubMed] [Google Scholar]
  • 605.Pelit A, Bagis T, Kayaselcuk F, Dursun D, Akova Y, Aydin P. Tear function tests and conjunctival impression cytology before and after hormone replacement therapy in postmenopausal women. Eur J Ophthalmol. 2003;13:337–42. doi: 10.1177/112067210301300402. [DOI] [PubMed] [Google Scholar]
  • 606.Schaumberg DA, Buring JE, Sullivan DA, Dana MR. Hormone replacement therapy and dry eye syndrome. JAMA. 2001;286:2114–9. doi: 10.1001/jama.286.17.2114. [DOI] [PubMed] [Google Scholar]
  • 607.Sullivan DA, Rocha EM, Aragona P, Clayton JA, Ding J, Golebiowski B, et al. TFOS DEWS II Sex, Gender, and Hormones report. Ocul Surf. 2017;15:284–333. doi: 10.1016/j.jtos.2017.04.001. [DOI] [PubMed] [Google Scholar]
  • 608.Xu N, Huang D, Yang H, Lai Z, Luo Q. Ocular surface characteristics and impression cytology in patients with active versus inactive Thyroid Eye Disease. Eye Sci. 2012;27:64–8. doi: 10.3969/j.issn.1000-4432.2012.02.003. [DOI] [PubMed] [Google Scholar]
  • 609.Huang D, Xu N, Song Y, Wang P, Yang H. Inflammatory cytokine profiles in the tears of thyroid-associated ophthalmopathy. Graefes Arch Clin Exp Ophthalmol. 2012;250:619–25. doi: 10.1007/s00417-011-1863-x. [DOI] [PubMed] [Google Scholar]
  • 610.Gilbard JP, Farris RL. Ocular surface drying and tear film osmolarity in thyroid eye disease. Acta Ophthalmol (Copenh) 1983;61:108–16. doi: 10.1111/j.1755-3768.1983.tb01401.x. [DOI] [PubMed] [Google Scholar]
  • 611.Iskeleli G, Karakoc Y, Abdula A. Tear film osmolarity in patients with thyroid ophthalmopathy. Jpn J Ophthalmol. 2008;52:323–6. doi: 10.1007/s10384-008-0545-7. [DOI] [PubMed] [Google Scholar]
  • 612.Ujhelyi B, Gogolak P, Erdei A, Nagy V, Balazs E, Rajnavolgyi E, et al. Graves’ orbitopathy results in profound changes in tear composition: a study of plasminogen activator inhibitor-1 and seven cytokines. Thyroid. 2012;22:407–14. doi: 10.1089/thy.2011.0248. [DOI] [PubMed] [Google Scholar]
  • 613.Yazici A, Sari ES, Sahin G, Kilic A, Cakmak H, Ayar O, et al. Change in tear film characteristics in visual display terminal users. Eur J Ophthalmol. 2015;25:85–9. doi: 10.5301/ejo.5000525. [DOI] [PubMed] [Google Scholar]
  • 614.Kojima T, Ibrahim OM, Wakamatsu T, Tsuyama A, Ogawa J, Matsumoto Y, et al. The impact of contact lens wear and visual display terminal work on ocular surface and tear functions in office workers. Am J Ophthalmol. 2011;152:933–40. doi: 10.1016/j.ajo.2011.05.025. [DOI] [PubMed] [Google Scholar]
  • 615.Jansen ME, Begley CG, Himebaugh NH, Port NL. Effect of contact lens wear and a near task on tear film break-up. Optom Vis Sci. 2010;87:350–7. doi: 10.1097/OPX.0b013e3181d951df. [DOI] [PubMed] [Google Scholar]
  • 616.Schlote T, Kadner G, Freudenthaler N. Marked reduction and distinct patterns of eye blinking in patients with moderately dry eyes during video display terminal use. Graefes Arch Clin Exp Ophthalmol. 2004;242:306–12. doi: 10.1007/s00417-003-0845-z. [DOI] [PubMed] [Google Scholar]
  • 617.Freudenthaler N, Neuf H, Kadner G, Schlote T. Characteristics of spontaneous eyeblink activity during video display terminal use in healthy volunteers. Graefes Arch Clin Exp Ophthalmol. 2003;241:914–20. doi: 10.1007/s00417-003-0786-6. [DOI] [PubMed] [Google Scholar]
  • 618.Patel S, Henderson R, Bradley L, Galloway B, Hunter L. Effect of visual display unit use on blink rate and tear stability. Optom Vis Sci. 1991;68:888–92. doi: 10.1097/00006324-199111000-00010. [DOI] [PubMed] [Google Scholar]
  • 619.Fenga C, Aragona P, Cacciola A, Spinella R, Di Nola C, Ferreri F, et al. Meibomian gland dysfunction and ocular discomfort in video display terminal workers. Eye (Lond) 2008;22:91–5. doi: 10.1038/sj.eye.6703025. [DOI] [PubMed] [Google Scholar]
  • 620.Fenga C, Aragona P, Di Nola C, Spinella R. Comparison of ocular surface disease index and tear osmolarity as markers of ocular surface dysfunction in video terminal display workers. Am J Ophthalmol. 2014;158:41–8. doi: 10.1016/j.ajo.2014.03.007. [DOI] [PubMed] [Google Scholar]
  • 621.Cardona G, Garcia C, Seres C, Vilaseca M, Gispets J. Blink rate, blink amplitude, and tear film integrity during dynamic visual display terminal tasks. Curr Eye Res. 2011;36:190–7. doi: 10.3109/02713683.2010.544442. [DOI] [PubMed] [Google Scholar]
  • 622.Abusharha AA, Pearce EI, Fagehi R. Effect of Ambient Temperature on the Human Tear Film. Eye Contact Lens. 2016;42:308–12. doi: 10.1097/ICL.0000000000000210. [DOI] [PubMed] [Google Scholar]
  • 623.Torricelli AA, Novaes P, Matsuda M, Alves MR, Monteiro ML. Ocular surface adverse effects of ambient levels of air pollution. Arq Bras Oftalmol. 2011;74:377–81. doi: 10.1590/s0004-27492011000500016. [DOI] [PubMed] [Google Scholar]
  • 624.Torricelli AA, Matsuda M, Novaes P, Braga AL, Saldiva PH, Alves MR, et al. Effects of ambient levels of traffic-derived air pollution on the ocular surface: analysis of symptoms, conjunctival goblet cell count and mucin 5AC gene expression. Environ Res. 2014;131:59–63. doi: 10.1016/j.envres.2014.02.014. [DOI] [PubMed] [Google Scholar]
  • 625.Gomes JAP, Azar DT, Baudouin C, Efron N, Hirayama M, Horwath-Winter J, et al. TFOS DEWS II Iatrogenic report. Ocul Surf. 2017;15:511–38. doi: 10.1016/j.jtos.2017.05.004. [DOI] [PubMed] [Google Scholar]
  • 626.Jordan A, Baum J. Basic tear flow - does it exist? Ophthalmology. 1980;87:920–30. doi: 10.1016/s0161-6420(80)35143-9. [DOI] [PubMed] [Google Scholar]

RESOURCES